You are on page 1of 23

TYPE Review

PUBLISHED 21 November 2023


DOI 10.3389/feart.2023.1159412

The distribution of volatile


OPEN ACCESS elements during rocky planet
EDITED BY
Mathieu Roskosz,
Muséum National d’Histoire Naturelle,
formation
France

REVIEWED BY Terry-Ann Suer 1,2*, Colin Jackson 3, Damanveer S. Grewal 4,


Sung Keun Lee,
Seoul National University, Republic of
Celia Dalou 5 and Tim Lichtenberg 6
Korea 1
Laboratory for Laser Energetics, University of Rochester, Rochester, NY, United States, 2Steward
Sami Mikhail, Observatory, University of Arizona, Tucson, AZ, United States, 3Earth and Environmental Sciences, Tulane
University of St Andrews, United Kingdom University, New Orleans, LA, United States, 4Division of Geological and Planetary Sciences, California
*CORRESPONDENCE Institute of Technology, Pasadena, CA, United States, 5Université de Lorraine, CNRS, Centre de
Terry-Ann Suer, Recherches Pétrographiques et Géochimiques, Nancy, France, 6Kapteyn Astronomical Institute,
tsuer@lle.rochester.edu University of Groningen, Groningen, Netherlands

RECEIVED 05 February 2023


ACCEPTED 02 November 2023
PUBLISHED 21 November 2023

CITATION
Core segregation and atmosphere formation are two of the major processes that
Suer T-A, Jackson C, Grewal DS, Dalou C redistribute the volatile elements—hydrogen (H), carbon (C), nitrogen (N), and sulfur
and Lichtenberg T (2023), The (S)—in and around rocky planets during their formation. The volatile elements by
distribution of volatile elements during
rocky planet formation. definition accumulate in gaseous reservoirs and form atmospheres. However, under
Front. Earth Sci. 11:1159412. conditions of early planet formation, these elements can also behave as siderophiles
doi: 10.3389/feart.2023.1159412 (i.e., iron-loving) and become concentrated in core-forming metals. Current
COPYRIGHT models of core formation suggest that metal-silicate reactions occurred over a
© 2023 Suer, Jackson, Grewal, Dalou and wide pressure, temperature, and compositional space to ultimately impose the
Lichtenberg. This is an open-access
article distributed under the terms of the chemistries of the cores and silicate portions of rocky planets. Additionally, the
Creative Commons Attribution License solubilities of volatile elements in magmas determine their transfer between the
(CC BY). The use, distribution or planetary interiors and atmospheres, which has recently come into sharper focus in
reproduction in other forums is
permitted, provided the original author(s) the context of highly irradiated, potentially molten exoplanets. Recently, there has
and the copyright owner(s) are credited been a significant push to experimentally investigate the metal-silicate and magma-
and that the original publication in this gas exchange coefficients for volatile elements over a wide range of conditions
journal is cited, in accordance with
accepted academic practice. No use, relevant to rocky planet formation. Qualitatively, results from the metal-silicate
distribution or reproduction is permitted partitioning studies suggest that cores of rocky planets could be major reservoirs of
which does not comply with these terms. the volatile elements though significant amounts will remain in mantles. Results
from solubility studies imply that under oxidizing conditions, most H and S are
sequestered in the magma ocean, while most N is outgassed to the atmosphere,
and C is nearly equally distributed between the atmosphere and the interior. Under
reducing conditions, nearly all N dissolves in the magma ocean, the atmosphere
becomes the dominant C reservoir, while H becomes more equally distributed
between the interior and the atmosphere, and S remains dominantly in the interior.
These chemical trends bear numerous implications for the chemical differentiation
of rocky planets and the formation and longevity of secondary atmospheres in the
early Solar System and exoplanetary systems. Further experimental and modeling
efforts are required to understand the potential of chemical and physical
disequilibria during core formation and magma ocean crystallization and to
constrain the distributions of volatile elements in the interiors and atmospheres
of rocky planets through their formation and long-term geologic evolution.

KEYWORDS

planet formation, metal-silicate partitioning, magma-atmosphere exchange, core-


mantle differentiation, volatile elements, atmosphere formation, super-earth exoplanets

Frontiers in Earth Science 01 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

1 Introduction partitioning during core-mantle-differentiation indicated that the


mantle’s refractory siderophile abundances require metal-silicate
Hydrogen, carbon, nitrogen, and sulfur are abundant volatiles in partitioning at pressures from 25 to 30 GPa, equivalent to a magma
the atmospheres of terrestrial planets in the Solar System and are ocean depth of 800 km (Li and Agee, 2001). However, later
expected to similarly dominate the gaseous envelopes of extrasolar experiments found that average conditions of 45–55 GPa
rocky planets (Wordsworth and Kreidberg, 2022). Depending on the (~1,500 km depth) and ≥3500 K (Bouhifd and Jephcoat, 2011;
prevailing thermodynamic and compositional properties, these Siebert, et al., 2012; Fischer, et al., 2015) could better explain the
elements can be present in atmospheres in oxidized (e.g., H2O, mantle’s Ni and Co abundances. Large impacts may have provided
SO2, CO2) or reduced forms (e.g., H2, H2S, CO) (Gaillard, et al., the energy for such deep magma oceans and their associated
2022). A significant proportion of a planet’s volatile elements can be episodes of core formation (Tonks and Melosh, 1993; Davies,
locked up in deep reservoirs such as cores and mantles whose bulk et al., 2020), although rapid pebble accretion (Johansen, et al.,
compositions are largely determined during early planet formation 2021) could also have led to high temperatures and mantle melting.
(Rubie, et al., 2007). Atmospheric outgassing (magma-gas exchange) Recent experimental studies done on different groups of
can occur simultaneously with core-mantle differentiation while elements including highly siderophiles and volatiles have
planets are hot enough to sustain magma oceans (Elkins-Tanton, indicated that one set of P-T-f O2-X condition cannot account for
2012; Hirschmann, 2012). Prior to the outgassing of their secondary all the relevant geochemical observables such as the bulk silicate
atmospheres, large rocky planets (>0.2 ME) that formed early likely Earth (BSE) contents in highly siderophile and volatile elements.
accreted primary H and/or He nebular atmospheres that would have More integrated approaches have been developed which consider
interacted with their early magma oceans (Sasaki, 1990; Zahnle, the dispersal of impacting materials and degree of chemical
et al., 2010). Although Earth lost its primordial atmosphere (Catling, equilibration before segregating into a core and a mantle (Rubie,
et al., 2001), studies of exoplanetary systems have revealed that some et al., 2003; Rubie, et al., 2007; Deguen, et al., 2014). These multi-
rocky planets can retain such envelopes throughout their evolution stage approaches can account for some geochemical observables
(Misener and Schlichting, 2021). Atmospheres can also be strongly such as the moderately siderophile elemental abundances in the
modified by impact-induced shock heating and erosion (Matsui and mantles of smaller planetary bodies, which required metal-silicate
Abe, 1986; Zahnle, et al., 1988; Davies, et al., 2020; Zahnle, et al., reactions at lower pressure-temperature conditions (Righter and
2020). However, from a geoscientific perspective, the two main Drake, 1997; Steenstra, et al., 2016). Though core formation in
volatile redistribution mechanisms are core formation and magma terrestrial planets is generally thought to occur at pressures and
outgassing, which we discuss here with particular emphasis on temperatures where metal and silicate phases are largely immiscible,
recent measurements. Volatiles retained within rocky planets’ other large-scale processes such as the segregation of a sulfide matte
mantles can reside as various mineral or fluid components and may accompany or follow core segregation (Rubie, et al., 2016;
participate in geological and biochemical processes (Dasgupta and Steenstra, et al., 2020). Additionally, density estimates for rocky
Hirschmann, 2010; Li, et al., 2013; Mikhail and Sverjensky, 2014; exoplanets (Otegi, et al., 2020) and the bulk chemical abundances of
Armstrong, et al., 2015; Hirschmann, 2018). Significant amounts of disintegrated planetesimals in the atmospheres of polluted white
volatile elements, along with other light elements (e.g., O and Si), are dwarfs (Bonsor, et al., 2020; Bonsor, et al., 2022) indicate that the
also inferred to be present in the cores of Earth and other rocky process of planetary differentiation could be common in planetary
planets such a Mars, where they lower the melting temperature and systems.
density of iron, leading to partially to fully molten cores capable of The failure of single-stage homogenous models to find a unique
sustaining a dynamo (Badro, et al., 2014; Nimmo and Schubert, P-T-f O2-X condition that can explain all geochemical observables
2015; Brennan, et al., 2020; Lv and Liu, 2022). suggests not only that Earth formed in stages and experienced
The Earth’s core and mantle inherited their elemental disequilibria, but also that accreting material might have changed
abundances during Earth’s differentiation. Core-mantle over time (Wade, et al., 2012). Indeed, N-body simulations of
segregation involved chemical reactions between Fe and silicates planetary accretion (Chambers and Wetherill, 1998; O’Brien,
in a reduced (Fe-alloy saturated) magma ocean (Wade and Wood, et al., 2014) imply that fully formed rocky bodies experience
2005; Wood, et al., 2006). Geochemical evidence—the abundance of many large impacts during their growth and that radial mixing
refractory siderophile elements such as Ni, Co, W and Mo in mantle of inner and outer Solar System materials occurs concomitantly.
peridotite—coupled with metal-silicate partitioning experiments This heterogeneous framework allows for many episodes of core
have led to inferences about the average conditions of core formation to occur at varying pressures, temperatures, and
formation (Ringwood, 1977). The earliest framework for compositions (Rubie, et al., 2003; Albarede, 2009; Schönbächler,
understanding core-mantle differentiation was provided by et al., 2010). Crucially, heterogeneous models driven by the mass
single-stage models in which metals accumulated at the bottom delivery scenarios of N-body simulations tend to favor the delivery
of the magma ocean before descending into the core (Stevenson, of more volatile-rich materials, ultimately sourced from the outer
1981; Stevenson, 1988; Karato and Murthy, 1997; Solomatov, 2007). Solar System, later in the accretion process when the potential for
These models assumed that the entire core and magma ocean extreme P-T metal-silicate reactions is at its maximum (Halliday,
equilibrated at a common pressure-temperature-oxygen fugacity 2013; O’Brien, et al., 2014; Rubie, et al., 2016). However, analyses of
and composition (P-T-f O2-X) (Jones and Drake, 1986) and led small rocky bodies and coupled to astrophysical and geochemical
to valuable insights on how redox state and planetary size affect core models suggest that volatile elements were present in the inner Solar
and mantle chemistry (Righter and Drake, 1996; Wade and Wood, System well before the later stages of planetary formation (Bar-Nun
2005; Corgne, et al., 2009). Early experimental studies of elemental and Owen, 1998; Busemann, et al., 2006; Marty, 2012; Halliday,

Frontiers in Earth Science 02 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

2013; Alexander, 2017; McCubbin and Barnes, 2019; Grewal, et al., rocky planets in these exoplanetary systems (Lichtenberg, et al.,
2021c; Deligny, et al., 2021; Grewal, 2022; Grewal and Asimow, 2019; Kite, et al., 2020; Lichtenberg, et al., 2021b; Dorn and
2023). Lichtenberg, 2021; Gaillard, et al., 2021; Kite and Schaefer, 2021;
Another consequence of heterogeneous accretion models is the Lichtenberg and Krijt, 2021; Schlichting and Young, 2022; Wolf,
implication that rocky planets can experience several magma ocean et al., 2022). Laboratory parameterizations of volatile partitioning
stages throughout their evolution (Elkins-Tanton, 2012). Magma- and solubilities are increasingly being incorporated into the models
gas interactions during magma ocean stages lead to volatile which calls for an effort to contextualize the measurements. In this
outgassing at temperature and compositional conditions related, review, we aim to do just that, such that the synergy between
but not equal, to those inferred for core formation (Zahnle, et al., laboratory investigations and astronomical characterizations of
2010; Hirschmann, 2016; Dasgupta and Grewal, 2019; Lichtenberg, hot exoplanets can be maximized.
2022). Volatiles are, by their nature, prone to partitioning into the
gas; accordingly, an increasing effort is being invested into
developing models to account for the atmospheres of accreting 2 Metal-silicate partitioning
planets as large volatile reservoirs, effectively limiting
incorporation into magma oceans or partitioning into cores 2.1 Parameterizing experiments
during planetary accretion. Factors such as oxygen fugacity and
magma ocean depth and volatile content strongly affect the resulting In addition to gravitational segregation, core-mantle
atmosphere and residual mantle compositions (Hirschmann, 2016; differentiation in terrestrial planets can be thought of in terms of
Grewal, et al., 2021b; Gaillard, et al., 2021; Jackson et al., 2021). the chemical separation of core Fe from mantle oxides and silicates.
Models are now being developed to account for the distribution of Thus, laboratory experiments study the chemical transfer of relevant
volatile elements into atmospheres with varying thermodynamic species between metal alloys and silicate/oxide mixtures. The
and compositional properties (Salvador, et al., 2017; Nikolaou, et al., parameterization of metal-silicate partitioning reactions and
2019; Lichtenberg, et al., 2021b; Bower, et al., 2022). Although experimental results has been covered extensively in several
solubility establishes the equilibrium distribution of a volatile recent works (Wood, et al., 2006; Corgne, et al., 2008a;
element between a primordial atmosphere and its underlying Blanchard, et al., 2017; Huang and Badro, 2018; Chidester, et al.,
magma ocean, increasing attention is being paid to how kinetic 2022), and only a brief overview is given here.
processes, including diffusion and convective stirring within The equilibrium partitioning of an element between metal and
planetary interiors, may affect the outgassing and ingassing of silicate can be thought of as an exchange reaction between molten
magma oceans (Salvador and Samuel, 2023). metal (M) and oxide:
Partition coefficients and solubilities are the fundamental data n n
needed to parameterize core-mantle differentiation and magma MO n2 + Fe <  > FeO+M (1)
2 2
ocean-atmosphere exchange models. These parameters have been
incorporated into models that are used to understand how volatile where n is the valence of element M. At equilibrium, the partition
elements are fractionated and cycled among major planetary coefficient D is the ratio of the concentration (here in terms of molar
reservoirs—i.e., cores, mantles, and atmospheres—for a range of fraction X, but can also be in terms of weight percent, wt. %) of an
bodies, including Earth, Mars, Venus, and differentiated element in the metal phase to that in the silicate phase:
planetesimals such as the parent bodies of iron meteorites XM
DM  (2)
(Hirschmann, 2012; Wordsworth, 2016; Gaillard, et al., 2021; XMOn/2
Jackson et al., 2021; Li, et al., 2021; Grewal, et al., 2022a;
Lichtenberg, et al., 2022; Grewal and Asimow, 2023). Although The partition coefficient is related to exchange coefficient of the
early models were tuned and benchmarked to the Earth, the reaction, KD, by normalizing by the partition coefficient for Fe:
parameterization made available by recent partitioning and DM
solubility measurements allows the models to be adapted to a KD (M)  (3)
Dn/2
Fe
wide range of other planetary formation scenarios. Thousands of
rocky exoplanets have been discovered by surveys such as the Kepler That is, KD is the ratio of the partition coefficient of element M to
Mission (Batalha, 2014) and the Transiting Exoplanet Survey that of iron and is independent of the Fe concentrations during the
Satellite (Kaltenegger, et al., 2019), and atmospheres of super- reaction.
Earths and sub-Neptunes are now being characterized by the KD is related to the equilibrium constant K of reaction (1), which
James Webb Space Telescope (JWST) (Mansfield, et al., 2019; can be defined in terms of KD and the activity coefficient (γ) of the
Ding and Wordsworth, 2022). In particular, in its first year of metal and silicate phases (E.g., Chidester, et al., 2022):
observations, JWST has demonstrated its capabilities to constrain
silicate n/2
the presence or absence of large volatile envelopes around short- γmetal γFeO 
log K(M)  log KD (M) + log M
n/2 + log silicate
period rocky exoplanets (Greene, et al., 2023; Ih, et al., 2023; Moran, metal
γFe  γMOn/2
et al., 2023; Zieba, et al., 2023). After reasonable observational times
b P
over the next few observational cycles, JWST will be able to constrain a+ +c (4)
T T
models of the volatile contents and redox states of individual rocky
exoplanets (Kempton, et al., 2023; Piette, et al., 2023). It is therefore Note that the oxygen fugacity of the experiments can be
crucial to continue efforts to model the formation and evolution of expressed in terms of the Fe–FeO equilibrium: Fealloy + 1/2 O2 =

Frontiers in Earth Science 03 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FeOsilicate melt, i.e., the Iron-Wüstite (IW) redox buffer (E.g., Corgne, d dM
((1 − F)Mxm )[(1 − F)xme + kF(xce − DX xm )] (10)
et al., 2008a; Campbell, et al., 2009; Chidester, et al., 2022): dt dt
aFe γ XFe d dM
ΔIW −2log10  −2log10 Fe (5) (FMxc )[kFDX xm + (1 − k) Fxce ] (11)
aFeO γFeO XFeO dt dt
where xm and xc are the concentrations of species x in the mantle and
Where a in this case refers to the activity.
core respectively of the resultant body. M is the mass of the Earth
For reaction (1), K is related to the Gibbs free energy of the
which changes with time t. Dc evolves with P-T-f O2-X conditions as
reaction (E.g., Chapter 11 of Ganguly, 2008):
the planet accretes. This framework can be adapted to different
ΔGo ΔHo − TΔSo + PΔVo evolution and accretion paths, including accreting bodies of
log K(M) −  (6)
RT RT different sizes and compositions and varying degrees of metal-
silicate equilibration between impactors and the magma ocean.
where R is the gas constant, T temperature, P pressure, and ΔHo,
The first term on the right-hand side of Eq. 10 represents
ΔSo, and ΔVo are the changes in enthalpy, entropy, and volume of
material from the embryo’s mantle that is added to the Earth’s
reaction (1), respectively. The natural log of the partition coefficient,
mantle, whereas the second term represents the mantle that re-
log D(M) can then be expressed in terms of the above variables and
equilibrated with the impactor core. Partial equilibration due to
constants:
inefficient mixing between large impactors towards the end of
b P ΔIW γmetal
log D(M)  a + +c −d − log M
(7) accretion (Rubie, et al., 2003; Dahl and Stevenson, 2010) is
T T 2 metal n/2
γ  Fe introduced by the parameter k, which is the fraction of the
embryo’s core that equilibrates with the planet’s mantle. The
The constants a, b, c, and d are determined using multivariate expressions on the right-hand side of Eq. 11 represent the two
least-squares linear regression of the measured Ds, Ps, Ts, and paths that the metal can take to the core: (1 − k) Fcce is the fraction of
compositions. The parameterization can also include activity the embryo’s core that is added directly to the planet’s core whereas
terms for the light components in the alloy and silicate melt kFDxcm is the fraction that equilibrates with the mantle.
polymerization. The expression obtained from the regression can Alternatively, partial equilibration can be implemented using a
then be used to determine partition coefficients over a wide turbulent mixing parameterization derived from fluid dynamic
P-T-f O2-X space, although care must be taken when regressing experiments (Deguen, et al., 2014), which is identical in terms of
because a, b, and c may significantly change with variations in mass conservation to the approach of Rudge et al. (2010), but the
experimental conditions (Walter and Cottrell, 2013). By coupling single fraction k is replaced by an efficiency factor ξ:
these functions to core formation models, we can calculate the re-
distribution of elements between cores and mantles during k
ξ (12)
differentiation. 1 + DΔX

where Δ is a metal dilution ratio that is related to the relative


2.2 Core formation models densities of the equilibrating metal and silicate (Deguen et al., 2014).

In the single-stage core formation model, a planet is assumed to


differentiate instantaneously and the re-distribution of a chemical 2.3 Metal-silicate partitioning behaviors of
species between core and mantle is determined at one set of the volatile elements
conditions. Using the formalism of (Rudge, et al., 2010), the
mass of species c in the body is conserved as follows: The effect of core formation on the distribution of volatile
elements has primarily been investigated through metal-silicate
xb  Fxce + (1 − F)xme (8)
partitioning experiments at conditions relevant to small
where xb is the bulk composition, xme is the concentration in the protoplanets and terrestrial planet embryos (0.5 GPa ≤ p ≤
mantle embryo, xce is the concentration in the core of the embryo, 20 GPa and T up to 2500 K) in large-volume press studies
and F is the body’s core mass fraction, which varies in the Solar (Dasgupta, et al., 2013; Roskosz, et al., 2013; Boujibar, et al.,
System from 0.25 (Mars) to 0.8 (Mercury). The effective partitioning 2014; Chi, et al., 2014; Stanley, et al., 2014; Li, et al., 2015; Li,
Dx in the embryo is defined at one set of P-T-X conditions and et al., 2016a; Li, et al., 2016b; Dalou, et al., 2017; Clesi, et al., 2018;
assumed constant in single-stage models: Speelmanns, et al., 2018; Tsuno, et al., 2018; Grewal, et al., 2019a;
xce Grewal, et al., 2019b; Dalou, et al., 2019; Kuwahara, et al., 2019;
Dx  (9) Malavergne, et al., 2019; Grewal, et al., 2021a; Grewal, et al., 2021b;
xme
Fichtner, et al., 2021; Jackson et al., 2021; Kuwahara, et al., 2021;
In multi-stage core formation models, partition coefficients Zhang and Li, 2021; Grewal, et al., 2022a; Grewal, et al., 2022b; Shi,
evolve with the magma ocean depth to account for multiple or et al., 2022; Grewal and Asimow, 2023). More recently, experiments
continuous differentiation events (Rubie, et al., 2003). The performed in laser-heated diamond anvil cells (DACs) have
conservation of mass of species x into the accreting planet’s core extended the P-T range of measurements on volatile elements to
and mantle can integrated numerically from Eqs 9, 10 following the average conditions thought to be relevant to core formation in
from Rudge et al. (2010). Similar formalisms are also developed the Earth and beyond (20 GPa ≤ p ≤ 108 GPa, and T up to ~5500 K)
elsewhere (Rubie, et al., 2011): (Roskosz, et al., 2013; Suer, et al., 2017; Fischer, et al., 2020; Jackson

Frontiers in Earth Science 04 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

TABLE 1 Recommended values of volatile element partition coefficients (Dx) over a range of P-T-f O2-X conditions relevant to different magma ocean depth/
planetary size and redox conditions. Functional forms from the literature are used to calculate these partition coefficients, see Supplementary Appendix SA (H
(Malavergne, et al., 2019; Tagawa, et al., 2021), N (Grewal, et al., 2021b), C (Malavergne, et al., 2019; Fischer, et al., 2020; Blanchard, et al., 2022), S (Boujibar, et al.,
2014; Suer, et al., 2017). Note that these recommended values based on specific set of conditions and different values will be obtained if conditions/variables in the
functions are changed. Ds are rounded off to the nearest significant decimal. Uncertainties on the Ds are dependent on the uncertainties in the fits to the
functional forms and can be as large as ± 20%.

Planetesimals/shallow magma ocean Embryo/intermediate magma ocean Planet/deep magma ocean

Reduced Oxidized Reduced Oxidized Reduced Oxidized (ΔIW


(ΔIW ~ −4) (ΔIW ~ −1) (ΔIW ~ −4) (ΔIW ~ −1) (ΔIW ~ −4) ~ −1 to ~1.3)
P 5 5 20 20 60 60
(GPa)

T (K) 2000 2000 3,000 3,000 4,000 4,000

DS 3 60 30 183 10 173

DC 316 10 43 1.4 0.5 to 20 0.1 to 6

DH 0.3 0.07 60 1.9 100 30

DN 5 265 1 39 no data no data

et al., 2021; Tagawa, et al., 2021; Blanchard, et al., 2022; Chidester, compilation in Figure 1); it readily forms FeHx at pressures
et al., 2022). In the following subsections, we discuss these above 5 GPa, but decomposes at ambient conditions (Badding,
measurements and their major implications for the distribution et al., 1991; Okuchi, 1997). Okuchi (1997) indirectly (i.e., by
of volatile elements between the cores and mantles of the Earth and estimating the volume of gas bubbles) determined DH at 7.5 GPa
other rocky planets during their formation. We also provide and 1,200°C–1,500°C in large-volume press experiments to be ~1 on
recommended values of metal-silicate partition coefficients for a average and proposed that H could be a primary light element in
range of planetary formation scenarios (See Table 1 and details in Earth’s core. More recently, large-volume press experiments have
Supplementary Appendix SA). yielded new estimates for DH up to 20 GPa, as constrained by a
combination of analytical techniques including electron probe
2.3.1 Hydrogen microanalysis (EPMA), nuclear microprobe resonance (NMR),
Hydrogen is the most cosmochemically abundant of the and electron recoil detection analysis (ERDA) inferred from the
volatile elements and is found in significant concentrations in latter experiments (Clesi, et al., 2018; Malavergne, et al., 2019).
a wide range of meteorites (Alexander, 2019a; Alexander, 2019b; Resultant DH values from these two studies are lower than unity
McCubbin and Barnes, 2019; Lewis, et al., 2022; Peterson, et al., (lithophile behavior), but thermodynamic modeling also suggested
2023). The D/H ratios of various terrestrial, meteoritic, and that higher pressure conditions could make H more siderophile. It is
planetary reservoirs have been crucial in constraining the important to note that these studies measured sample compositions
origins of hydrogen on Earth, which measurements indicate after quenching, and it is possible that hydrogen diffused out of their
could be due to contributions from different reservoirs quench products, leading to large biases on DH. Carbon saturation in
accreted at different times throughout the Solar System’s the metals of the latter experiments could also have contributed to
history (Albarede, et al., 2013; Halliday, 2013). Studies of lowering DH.
hydrogen metal-silicate partitioning behavior (and solubility, Measurements in laser-heated DAC experiments indirectly
discussed in section 3.1.1) have recently become important in inferred H content from the unit cell expansion with in situ X-ray
the context of sub-Neptune exoplanets, which are thought to diffraction (Tagawa, et al., 2021) and found H to be siderophile, with
have retained thick primodrial H and/or He envelopes (Fulton, DH > 29 at 30–60 GPa and 3,100–4600 K, in agreement with
et al., 2017; Lichtenberg, et al., 2021a; Dorn and Lichtenberg, theoretical calculations (Yuan and Steinle-Neumann, 2020). When
2021; Rogers and Owen, 2021; Schlichting and Young, 2022). coupled with a core formation model, those values imply that up to
There is also isotopic evidence that the Earth and Mars could 0.6 wt. % H could have been incorporated into the Earth’s core
have had an early H-rich atmosphere (Tian, et al., 2005; Sharp, and >0.15 wt. % into the cores of planets more massive than
2017; Pahlevan, et al., 2022). Hydrogen could also have been 0.1ME if water was accreted during the main stage of accretion.
ingassed into Earth’s magma ocean if it was in contact with such
an atmosphere, and may have ultimately partitioned into the core 2.3.2 Carbon
(Wu, et al., 2018; Young, et al., 2023). Likewise, the cores of The presence of carbon in Earth’s and other planetary cores has
H-rich exoplanets could also have inherited nebular H, which been inferred from its affinity for iron at ambient conditions and its
would have influenced their total water abundances (Kimura and depletion in the BSE relative to carbonaceous chondrites (CIs)
Ikoma, 2022). (Hirschmann, 2016). The results of metal-silicate partitioning
Though hypothesized to be a low-density core component, its experiments at pressures up to 15 GPa in large-volume presses
tendency to diffuse and exsolve makes measuring the metal-silicate (Dasgupta, et al., 2013; Chi, et al., 2014; Malavergne, et al., 2019)
partitioning of hydrogen difficult, and data remain sparse (See indicate that although carbon is a highly siderophile (DC up to 105)

Frontiers in Earth Science 05 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 1
Hydrogen partition coefficients (log DH) plotted as a function of oxygen fugacity, parameterized here in terms of the iron-wüstite (ΔIW) buffer (A) and
experimental temperature (B). Data sources: (Okuchi, 1997; Clesi, et al., 2018; Malavergne, et al., 2019; Tagawa, et al., 2021).

FIGURE 2
Carbon partition coefficients (log DC) plotted as a function of oxygen fugacity, parameterized here in terms of the iron-wüstite (ΔIW) buffer (A) and
experimental temperature (B). Data sources: (Dasgupta, et al., 2013; Chi, et al., 2014; Armstrong, et al., 2015; Li, et al., 2016b; Dalou, et al., 2017; Tsuno,
et al., 2018; Grewal, et al., 2019a; Kuwahara, et al., 2019; Malavergne, et al., 2019; Fischer, et al., 2020; Grewal, et al., 2021a; Fichtner, et al., 2021;
Blanchard, et al., 2022).

at lower pressures conditions (~1 GPa), it will partition significantly Jackson et al., 2021), and immiscible C- and S-rich layers could have
less into Fe as pressure increases (e.g., DC ≈ 10 at 15 GPa) (See been contemporaneous in the cores of some planetesimals (Corgne,
compilation in Figure 2). Models based on these results, suggest that et al., 2008b).
the cores of smaller planetesimals could retain wt. % levels of carbon, Experimental results at higher P-T conditions in laser-heated
whereas those of larger bodies would be less enriched (<1 wt. % C) DACs show that carbon has a lowered affinity for iron at the
while their mantles would become progressively more C-rich during conditions of a deep terrestrial magma ocean (Fischer, et al.,
accretion (Kuwahara, et al., 2021). Though DC can increase as fO2 2020; Blanchard, et al., 2022). When coupled with models of
decreases from ΔIW −1 to −3 (Malavergne, et al., 2019), under Earth’s core formation and accretion, these partition coefficients
highly reducing conditions, Si will partition strongly into iron, imply that less than 1 wt. % C could be present in the cores of Earth
limiting C dissolution in cores of very reduced bodies. This and similar-sized bodies. Complementary mantle compositions
chemical trend has led to the suggestion that the Ureilite parent could be hundreds of ppm, depending on the timing of carbon
body or Mercury could have become carbon-saturated, possibly accretion.
leading to a graphite-rich crust (Vander Kaaden and McCubbin,
2015; Keppler and Golabek, 2019; Steenstra and van Westrenen, 2.3.3 Nitrogen
2020). The presence of sulfur and nitrogen in core materials can also Nitrogen is the most abundant component of Earth’s
affect the partitioning of carbon into the Fe (Grewal, et al., 2019b; atmosphere and its presence in meteoritic material indicates that

Frontiers in Earth Science 06 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 3
Nitrogen partition coefficients (log DN) plotted as a function of oxygen fugacity, parameterized here in terms of the iron-wüstite (ΔIW) buffer (A) and
experimental temperature (B). Data sources: (Roskosz, et al., 2013; Li, et al., 2015; Dalou, et al., 2017; Speelmanns, et al., 2018; Grewal, et al., 2019a;
Grewal, et al., 2019b; Dalou, et al., 2019; Grewal, et al., 2021b; Jackson et al., 2021; Shi, et al., 2022).

it was present in early planetary building blocks (Marty, 2012; measurements at higher P-T conditions will be valuable to
Grewal, et al., 2021b; Grewal, 2022; Grewal and Asimow, 2023). further investigations of N incorporation into the cores of large
Its depletion in Earth’s mantle relative to C (Marty, 2012) suggests planets.
that N could have been incorporated into the Earth’s core.
Furthermore, it has been noted from early solubility experiments 2.3.4 Sulfur
that N is increasingly incorporated into iron with increasing Sulfur’s affinity for iron at ambient conditions and its presence
pressures of up to 7 GPa (Roskosz, et al., 2013; Speelmanns, in iron meteorites imply that it could be a component of
et al., 2018). More recent metal-silicate partitioning experiments protoplanets and planetary cores (Dreibus and Wanke, 1985;
in large-volume presses (Dalou, et al., 2017; Grewal, et al., 2019a; Chabot, 2004; Gounelle and Zolensky, 2014). Combined with its
Dalou, et al., 2019) found N to be siderophile (10 < DN < 31) at cosmochemical abundance, these factors have been used to argue
1–7 GPa and up to 1800°C (See Figure 3); its partitioning behavior that S is one of the light alloying components in Earth’s and other
being sensitive to P, T, f O2, and the presence of light elements such planetary cores. Its incorporation into Fe has been studied
as S in Fe (Grewal, et al., 2019a). The sensitivity of N partitioning to extensively in an attempt to explain the density deficit in Earths’s
oxygen fugacity suggests that it dissolves as nitride (N3−) in silicate core (i.e., relative to that of pure iron) and the properties of other
melts and as neutral N in core metals (Dalou et al., 2019; Grewal planetary cores (Fei, et al., 1995; Morard, et al., 2013; Boujibar, et al.,
et al., 2020), and this change in oxidation state makes N partitioning 2020; Brennan, et al., 2020). S is also known to affect the behaviors of
relatively sensitive to oxygen fugacity compared to the other volatile other chemical species in various geological settings, particularly
elements. chalcophile elements (Jana and Walker, 1997; Mahan, et al., 2018)
Nitrogen is poorly soluble in magmas at oxygen fugacities and C (Li, et al., 2016b; Tsuno, et al., 2018). Experiments in large-
around IW (Libourel et al., 2003). Thus, magmatically active volume presses up to 30 GPa (Li and Agee, 2001; Boujibar, et al.,
bodies can lose a large fraction of their N to their atmosphere. 2014); showed the increasing siderophile tendency of S with
Combined differentiation, outgassing, and accretion models show increasing pressure, lending support to its incorporation into the
that protoplanets that differentiated early could have been depleted cores of planetesimals and planetary embryos. For example, Mars,
in N (Grewal, et al., 2021b). However, if differentiation occurred which is the size of a planetary embryo and more volatile rich than
later at the embryonic stage, nitrogen reservoirs could have been Earth, could contain up to ~20 wt. % S in its core (Brennan, et al.,
maintained within cores while the mantles would have remained 2020) while it has been speculated that Mercury’s core could be
depleted, a scenario which could explain the N signature of the surrounded by an FeS layer (Namur, et al., 2016). Experiments in
terrestrial mantle (Grewal, et al., 2021c). Limited laser-heated DAC laser-heated DACs ~29–100 GPa, up to 5300 K (Suer, et al., 2017;
data demonstrate that N remains siderophile up to 26 GPa and Mahan, et al., 2018; Jackson et al., 2018; Chidester, et al., 2022);
~3500 K (Jackson et al., 2021), further suggesting that it could be found DS values that are an order of magnitude lower on average
sequestered into larger planetary cores. This behavior is qualitatively than large-volume press results (ranging from less than 10–100),
different from C and S, which both become less siderophile at suggesting that sulfur’s siderophile tendency does increase with
increased pressures and temperatures. This change in siderophile pressure, but is also strongly suppressed at high temperatures
behavior among the volatile elements has significant implications for (see compilation in Figure 4). When the results of laser-heated
the S-C-N reservoirs on planetary embryos and larger planets DAC experiments are included in fitting the functional form for log
(Jackson et al., 2021). The measurements compiled in Figure 3 DS, the entropy term is an order of magnitude larger than if the fit
highlight that DN increases with increasing f O2. Obtaining included only the large-volume pressure cell results. This is generally

Frontiers in Earth Science 07 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 4
Sulfur partition coefficients (log DS) plotted as a function of oxygen fugacity, parameterized here in terms of the iron-wüstite (ΔIW) buffer (A) and
experimental temperature (B). Data sources: (Li and Agee, 1996; Ohtani, et al., 1997; Wade and Wood, 2001; Chabot and Agee, 2003; Wood, 2008; Mann,
et al., 2009; Rose-Weston, et al., 2009; Boujibar, et al., 2014; Suer, et al., 2017; Mahan, et al., 2018; Jackson et al., 2021; Chidester, et al., 2022).

the case with laser-heated DAC studies of other elements. Redox and products of large-volume press experiments lead to difficulty in
the chemical speciation of S have also been shown to play an determining stable liquid compositions. Likewise, the origin of
important role in its partitioning behavior (Mavrogenes and metallic inclusions in the silicate portions of DAC run products
O’Neill, 1999; Chidester, et al., 2022). has been controversial. Figure 5 shows examples of a large volume
When the functional form for DS is incorporated into models of and laser-heated DAC experiment post quench. Large metal
Earth’s core formation (Boujibar, et al., 2014; Suer, et al., 2017), less inclusions are usually excluded from analyses of silicate melt
than 2 wt. % S can partition into the core, in agreement with compositions, but smaller nanoscale inclusions (nanonuggets) are
cosmochemical (McDonough and Sun, 1995) and geophysical difficult to avoid, particularly with EPMA which integrates over
constraints (Badro, et al., 2014). However, depending on the several to tens of cubic microns. To assess this issue, NanoSIMS has
initial bulk sulfur content and/or the timing of sulfur accretion been used to obtain highly resolved measurements for S and C (Suer,
and the efficiency of core-mantle equilibration, more sulfur could be et al., 2017; Fischer, et al., 2020; Blanchard, et al., 2022). TEM
incorporated into planetary cores. Predicted sulfur compositions of measurements have also been performed to help resolve this
the Earth’s mantle could range from a few hundred to thousands of controversy in recent studies (Fischer, et al., 2015; Suer, et al.,
ppm also depending on the timing of sulfur accretion (Rubie, et al., 2017; Suer, et al., 2021), which suggest that nanoparticles form upon
2016; Suer, et al., 2017), overlapping with geochemical observations rapid quench but were dissolved during equilibrium melting.
suggesting that the BSE contains ~200 ppm S (McDonough and Sun, However, if nano-metal inclusions are contaminants, they could
1995; Wang and Becker, 2013). It is possible that magma ocean lower measured partition coefficients for the volatiles and other
crystallization leads to an increase in S concentrations (Rubie, et al., elements, significantly impacting the results of models that utilize
2016). Above the sulfur capacity at sulfide saturation (SCSS) limit, these measurements.
precipitation and segregation of a sulfide-rich matte could occur on De-volatilization and the loss of volatile species during
a planet-wide scale in both small and large bodies (O’Neill and quenching can lead to large uncertainties on experimental
Mavrogenes, 2002; Steenstra, et al., 2020), leading to a late pulse of measurements. The degree to which a high pressure-high
core formation if the sulfide phase can mobilize to the core-mantle temperature assemblage is preserved depends on the quench rate
boundary. of the experimental apparatus: example quench rates are 175°C/s in
piston-cylinder experiments, ~800°C/s in multi-anvil experiments,
and thousands of degrees per microsecond in DAC experiments.
2.4 Experimental limitations Magnesium-rich silicates (analogs of mafic terrestrial magma
oceans) are difficult to quench to a homogenous glass, and some
In the last decade, advances in experimental and analytical works have thus used basaltic and andesitic silicate compositions
techniques have enabled high-quality measurements of metal- (Suer, et al., 2017). Quenching is particularly important for
silicate partitioning coefficients and the solubilities of volatile hydrogen, which is known to escape from silicate melts and
elements. Nonetheless, the interpretation of some of these metals upon recovery at ambient conditions (Okuchi, 1997).
measurements remain controversial. Therefore, it has been particularly difficult to obtain reliable
Although not specific to volatile-element partitioning, partitioning measurements for hydrogen. Recently, the
ambiguity surrounds the interpretation of quench textures in combination of EPMA, NMR and ERDA have been used
both large-volume press and DAC experimental samples. successfully to measure partition coefficients for H and C (Clesi,
Dendritic textures and overgrowths of quench rims in run et al., 2018; Malavergne, et al., 2019).

Frontiers in Earth Science 08 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 5
Scanning electron microscope images of quench products from metal-silicate partitioning experiments on a large-volume press (A) (Dalou, et al.,
2017) and laser-heated diamond anvil cell (B). Bright regions indicate material with higher atomic number such as Fe-Ni alloys. The quenched silicate
regions of both samples contain small inclusions. In the case of diamond anvil cell experiments, inclusions can be as small as the nanometer scale.

The large concentrations of volatile elements used in the iron planetary bodies continuously redistribute their volatile element
alloys of some experiments can also affect partition coefficients. budgets as they grow and undergo new periods of core formation
Such concentrations are typically much higher than those observed and mantle melting. In the previous section, our focus was on the
in nature and can be in violation of Henry’s law. Thus, the sequestration of volatile elements within planetary cores via metal
application of such partition coefficients in models may segregation in magma oceans. In this section, however, our focus is
introduce additional, and difficult to quantify, uncertainties to understand how the presence of an atmosphere can affect the
(Kuwahara, et al., 2019; Grewal, et al., 2021a; Grewal, et al., distribution of volatile elements during core formation, because any
2022b; Shi, et al., 2022). gaseous molecule present in an atmosphere subtracts from the
Statistical regressions of measurements are used to obtain budget available to metal-silicate partitioning.
constants for the functional description of partitioning behavior.
Differences have been noted between partitioning behaviors in
large-volume press and DAC experiments. These have been 3.1 Gas solubilities in magma oceans
attributed to changes in chemical behaviors in the different P-T
regimes probed by these two types of apparatuses. This effect is Solubility is perhaps the most central parameter governing the
perhaps best documented for S and can be seen in the different chemical interaction of a primordial atmosphere and its underlying
magnitudes of the regression constants obtained across different magma ocean. More specifically, solubility quantifies the
experimental techniques (Boujibar, et al., 2014; Suer, et al., 2017). relationship between the fugacity of a gas species and its
Thus, a single functional form might not be sufficient to fully corresponding concentration in a condensed phase. This can be
describe datasets spanning large P-T ranges. In addition, linear expressed as:
functional relationships might not fully describe the convolution X i  f αi Si (13)
of the IW redox buffer with the behaviors of other elements
involved, and further modeling efforts are needed to deconvolve where X i is the concentration of element i dissolved in the
these effects. For these reasons, it is not recommended that the condensed phase, f αi is the fugacity of element i in the gas (or
functional forms be extrapolated beyond the ranges of the fluid) raised to an exponent that relates to the stoichiometry of the
measurements on which they are based. dissolution reaction(s), and Si is the solubility of element i in the
condensed phase. Solubility depends on the temperature and
pressure of the reaction as well as the compositions of the
3 Volatile element solubilities and condensed and gaseous phases. In our example, the condensed
magma-atmosphere interactions phase is the magma. Fugacity can be further deconvolved as:
f i  γi Pi (14)
Volatile elements distribute themselves between metal, silicate,
and gas during planetary formation (Chao, et al., 2021; Lichtenberg, where γi and Pi are the fugacity coefficient and the partial pressure
et al., 2022). Metal reacts with magma before its segregation to the of element i, respectively. The high temperatures and moderate
core, whereas gas and magma continuously exchange volatiles at the pressures in atmospheres suggest that the gases present will only
atmosphere-magma ocean interface. Thus, it is expected that have moderate deviations from ideality, and fugacity coefficients are

Frontiers in Earth Science 09 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

often assumed to be unity, such that fugacity and partial pressure are
equal. Eqs 13, 14 reveal the overarching role of atmospheric pressure
on atmosphere-magma interactions, as it is partial pressures that
drive the dissolution of volatile species into the magma.
The stability of a gas species depends on the prevailing P-T-X
conditions within the gas phase. For example, the stability of CO
with respect to CO2 can be expressed as:
COg + 0.5O2 g  CO2 g (15)

Le Chatelier analysis indicates that more oxidized conditions


favor CO2 stability compared to CO. In isolation, this reaction also
predicts that CO2 will be favored under lower temperature and
higher pressure given that the reactants have higher entropy and
volume, but all major gas species should be considered together
when making specific predictions regarding how species
concentration is affected by changing P-T conditions for any
bulk gas composition.
The example above stresses the importance of oxygen fugacity in
determining the stability of gas species. We note, however, that it is
not precisely known how oxygen fugacity varies throughout a
growing planet. It is known that, at depth, magma oceans are
saturated in liquid Fe alloy and are therefore strongly reduced FIGURE 6
environments (well below the IW oxygen buffer, as inferred by Schematic contours of H, C, and N speciation determined by
the low residual FeO concentrations in the bulk silicate portions of Raman spectroscopy in basaltic glasses synthesized in piston-cylinder
experiments at 1–2 GPa and 1,400°C–1,600°C, shown as a function of
planets), but isochemical decompression (or compression) of a fO2 (relative to the IW buffer) and H content (as equivalent H2O,
magma can drive homogeneous reactions that cause auto- wt. %). Raman and FTIR measurements are from Hirschmann (2012),
reduction or auto-oxidation (Zhang, et al., 2018; Armstrong, Armstrong et al. (2015), Dalou et al. (2019, 2022), and Grewal et al.
(2020). Blank areas represent conditions at which literature data are
et al., 2019). Moreover, magma ocean crystallization should lead currently lacking. This figure demonstrates that H, C, and N on their
to fO2 variations given the different partitioning of ferric and ferrous own may be present as multiple species at a given fO2 or water
Fe between minerals and melts. This implies that the oxygen fugacity content, and that they combine to form different molecules in natural
and synthetic glasses. fO2 of Earth and Mars formation are indicated.
associated with metal-silicate reactions need not be the same as that Sulfur speciation is not shown because the effect of H on S speciation
associated with gas-magma reactions, particularly because the has not yet been determined at fO2 relevant to magma ocean
magma ocean may become increasingly chemically isolated form conditions.

the atmosphere as solidification progresses (Bower et al., 2022).


However, our goal is not to review this dynamic and the potential
resulting redox stratification within planetary magma oceans thermodynamic data from the NIST WebBook to calculate
(Hirschmann, 2012). Rather, we introduce this concept to stress equilibrium constants of reactions between gases.
that a wide range of oxygen fugacities are likely to impact volatile Water vapor can dissolve into magma as either OH or H2O
element solubilities associated with planetary formation. We (molecular) following the reactions:
correspondingly organize our review of gas solubilities below by H2 O g  H2 O (melt) (17)
element, and within each element by oxygen fugacity. We also −2 −
H2 O(melt) + O (melt)  2OH (melt) (18)
emphasize that our goal is not to provide an unabridged review
of volatile element solubilities in magmas, but we simply seek to Reactions 17 and 18 predict that low f H2O values should favor
highlight the basic controls on volatile element solubilities and to the dissolution of OH; however, with increasing OH concentration,
provide reference solubility values that may serve to accelerate the relatively more H2O will dissolve as molecular H2O. Indeed, water
development of magma ocean-atmosphere interaction models. solubility has been experimentally shown to scale with f H2O0.5 at
lower concentrations, but with f H2O at higher concentration. The
3.1.1 Hydrogen transition from OH- to H2O-dominated water solubility typically
When reacted with oxygen, hydrogen gas forms water as (similar occurs at the wt. % level (Stolper, 1982), which requires f H2O >
to Reaction 15): 100 bars, whereas terrestrial planets are typically estimated to
H2 g + 0.5O2 g  H2 Og (16) contain between 100 and 1,000 ppm H2O integrated over their
silicate reservoirs, atmospheres, and oceans (Peslier, 2010;
Relative to H2, gaseous water molecules have a wide stability Halliday, 2013). This implies that only modest H2O pressures
field under geological conditions, as demonstrated by the ratio of could have been present in primordial atmospheres (~1–10 bars
f H2O/f H2 plotted as a function of ΔIW at 2273 K (Figure 7A). The on average), unless large amounts of H were lost from the
crossover of f H2O/f H2 ratio occurs near IW and is nearly atmospheres (Catling, et al., 2001). Moreover, the products in
independent of temperature. We note that we use Reaction 18 are favored at higher temperatures (Nowak and

Frontiers in Earth Science 10 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Figure 7 plots predicted H2O solubilities from MAGMASAT for


a peridotitic liquid (BSE of (McDonough and Sun, 1995) as a
function of PH2O (assumed ideal) and temperature. Solubility
scales close to PH2O0.5 and is essentially independent of
temperature, although recent experimental data suggest that
water solubility decreases with temperature. Fitting all predictions
together yields the following equation:
H2 Owt.%  0.0575·PH2 O0.5158 (19)

with PH2O in bars. The fact that MAGMASAT predicts that H2O
solubility scales close to PH2O0.5 is consistent with OH dominating
solubility up to 0.6 wt. % H2O dissolved in peridotitic magma.
Such predictions require significant extrapolation because most
water solubility determinations are limited to basaltic (or move
evolved) compositions and typical magmatic temperatures
(<1673 K) and should therefore be treated with caution.
Nevertheless, it is clear that, relative to other gases (see following
subsections), water vapor remains highly soluble in ultramafic
magma, even under magma ocean conditions.
FIGURE 7 Support for MAGMASAT model prediction comes from recent
MAGMASAT predictions of H2O solubility in an ultramafic melt at results from a laser levitation apparatus study (Sossi, et al., 2023).
different temperatures. The curve is fit to the 2273 K data and is given
This apparatus enables quenching of ultramafic glasses reacted with
by Eq. 19.
controlled f H2O at temperatures that are directly applicable to
magma oceans. Current data collected in systems with f H2O/
f H2 > 1 yield a similar solubility relationship to that predicted above:
Behrens, 1995), and, given that magma oceans exist at higher H2 Owt.%  0.0605·PH2 O0.47 (20)
temperatures than modern magmatic systems, it is expected that
OH is a major component of water solubility in growing planets, assuming a molar absorption coefficient (ε3550) of 5.1 m2/mol.
even under relatively reducing conditions (Figure 6). Under more reducing conditions, water vapor is progressively
Hydroxyl forms chemical bonds when dissolved in silicate melts, destabilized to produce H2. Spectral evidence demonstrates that H2
lowering melt viscosity and implying that water reacts with bridging gas dissolves as molecular H2 in magmas (Luth, et al., 1987). As a
oxygens in Reaction 18 to depolymerize the melt network. Magma neutral, non-polar molecule, H2 is thought to dissolve into the ionic
ocean liquids are depolymerized, meaning that they have few porosity of the melt structure, following the well-documented
bridging oxygens to support OH dissolution. Studies of water behavior of noble gases (Carroll and Stolper, 1993). The radius of
solubility in depolymerized liquids highlight that water also H2 is similar to that of helium, and indeed, experiments demonstrate
bonds with non-bridging oxygens, preferentially forming that H2 solubility is also similar to that of helium, at least for basaltic
complexes with higher field strength cations (Mysen and Virgo, melts (Hirschmann, 2012). The solubility of helium, or hydrogen,
1986; Xue and Kanzaki, 2004). It is therefore expected that water has not be experimentally investigated in peridotitic liquids, but
should dominantly dissolve as OH at the low fH2O conditions their solubilities can be estimated by extrapolating the relationship
relevant to magma ocean liquids. A recent work suggests that of ionic porosity and solubility for helium, assuming that hydrogen
peridotitic magmas contain relatively more OH than H2O solubility continues to track with helium towards more
compared to more polymerized magmas at equal total H2O depolymerized compositions on a molar basis. To do this, we
contents, such that H2O solubility in peridotite liquids will apply the algorithm developed for helium solubility in silicate
remain dominated by OH up to 5 wt. % H2O (Bondar, et al., 2023). melt by (Iacono-Marziano, et al., 2010) to estimate H2 solubility
Experimental determinations of water solubility in magma at 1 bar and 2273 K. Taking a melt composition equal to the BSE
oceans have historically been precluded by difficulties associated yields the relationship:
with working in ultramafic systems, although recent advances in H2wt.%  4.19 × 10−5 · PH2 (21)
rapid-quench multi-anvil and laser levitation techniques are highly
promising (Bondar, et al., 2023; Sossi, et al., 2023). In the absence of Experiments demonstrate that temperature has a modest effect
well-established data, we rely on models that allow for the P-T-X on the solubilities of noble gases (Jambon, et al., 1986) although data
dependencies of water solubility determined within the realm of are limited to relatively cool conditions compared to those
quenchable liquids to be extrapolated to the P-T-X realm of magma associated with magma oceans (<1873 K). It has been suggested
oceans. Several models have been developed (Newman and that higher temperatures should favor the dissolution of H2 into
Lowenstern, 2002; Papale, et al., 2006; Iacovino, et al., 2021), and magmas (Chachan and Stevenson, 2018; Kite, et al., 2020), and this
we use MAGMASAT (Ghiorso and Gualda, 2015) for our example effect could be significant given the wide temperature range over
because it is implemented in the popular MELTs algorithm and which magma oceans likely exist. The model of (Iacono-Marziano,
considers both OH and H2O dissolution. et al., 2010) predicts an order of magnitude increase in helium

Frontiers in Earth Science 11 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

solubility upon heating from 1773 to 3272 K, also suggesting the solved pressure of the atmosphere and that assumed to calculate K
importance of high temperatures in promoting H2 dissolution into until there is agreement.
magmas. We note that this approach can also be extended to predict core
We now apply the solubility laws for H to calculate its chemistry by adding a term that includes a partition coefficient (with
equilibrium distribution between a magma ocean and a implied P-T-X conditions for metal-silicate chemical exchange)
primordial atmosphere. Solving the equilibrium distribution for along with the available masses of silicate and metal to be
any volatile element is accomplished by solving a system of reacted to define the mass of the volatile element sequestered in
equations that include the solubility laws, statements of mass the core for the specified P-T-X conditions.
balance, and the equilibrium constants for gas phase equilibria to Figure 8 plots the distributions of H species between
define the relative stabilities of gas species as a function of magma atmospheres and magma oceans as a function of f O2 (relative to
ocean-atmosphere interface P-T-X conditions. Mass balance the IW oxygen buffer at 1 bar) and the total abundance of H. The
statements use total mass as a constraint, but the mass balance temperature dependance of H2 and H2O stability is largely canceled
constraint could also be the pressure of a gas species in the by referencing f O2 to the IW buffer. For simplicity, we assume a
atmosphere or the concentration of the relevant volatile element total atmospheric pressure of 100 bars and a temperature of 2773 K.
in the magma ocean. An assumed constant total pressure is supported by the relatively
To connect solubility laws to mass distribution, melt constant and high pressure that C species provide given the
concentration is scaled to mass by multiplying concentration by abundance of C in the BSE, as will be demonstrated in the next
the mass of the assumed magma ocean. Partial pressure (fugacity) is subsection.
scaled to atmospheric mass as: The crossover for equal partial pressures of H2 and H2O occurs
Px Ar near ΔIW−1 at 100 bars and 2273 K (Figure 8A). Water (steam) is
Mx  (22) much more soluble than H2, and this shifts the crossover point for
g
equal H2O and H2 concentrations in magma oceans to extremely
where Px is the partial pressure of gas x in the atmosphere, A is the reducing conditions (<ΔIW−6), such that H2O may dominate H
surface area of the planet, g is the gravitational constant, and r is the dissolution in magma oceans over a wide range of accretion
ratio of mass of the volatile element of interest in the gas molecule to scenarios despite the higher partial pressures of H2 in
the full mass of the gas molecule (e.g., 1 for H2 and 2/18 for H2O). It atmospheres (Figure 8B). In all scenarios, the magma ocean
is clear that the distribution of a volatile element depends on the size contains more H by mass than the atmosphere (Figure 8C). H
of the planet (A), its gravitational field (g), and the depth (or mass) of preferentially partitions into the magma ocean under oxidizing
the magma ocean, so these also must be specified. conditions, but as f O2 drops and f H2 rises, the distribution
We provide as an example the system of equations needed to becomes nearly equal, indicating the increased volatility of H in
solve the distribution of H. We apply a similar approach for other reduced systems. The small masses of H in primordial atmospheres
volatile elements in their respective sections. indicate that they have only a small capacity to limit the
incorporation of H into cores and determine the H abundances
K  PH2 Of O0.5
2 PH2  (23)
of the bulk silicate reservoirs (i.e., materials later derived from the
PH2 M MO PH2 O0.5158 M MO magma ocean-atmosphere system, that includes mantle, crust,
M H 4.19 × 10−5 · + 0.0575 ·
100 100 atmosphere, and oceans).
(24)
PH2 A PH2 OA182 
+ +
g g 3.1.2 Carbon
Under high f O2, CO2 is expected to be the dominant C-bearing
where K is the equilibrium constant for Reaction 24, MH is the total
gas species (Reaction 15). Spectroscopic results have identified that
mass of H in the atmosphere-magma ocean system, and MMO is the
C is present as carbonate groups in magmas when reacted with CO2
mass of the magma ocean. The equilibrium constant is defined by
gas (Mysen and Boettcher, 1975; Fine and Stolper, 1986), and this
the P-T conditions selected and the associated Gibbs energy change
observation suggests that the following set of reactions control CO2
for the reaction (calculated from data tabulated on the NIST
solubility in magmas:
WebBook and the ideal gas law). The factors of 100 are required
in the leftmost terms on the righthand side to convert from wt. % to CO2 g  CO2 (melt) (25)
−2
wt. fraction. The system is solved for the partial pressures of H2 and CO2 (melt) + O (melt)  CO−2
3 (melt) (26)
H2O (PH2 and PH2O). This assumes ideal gas behavior; high
temperature-moderate pressure systems (hundreds of bars) only From these reactions, C solubility should linearly scale with CO2
have small deviations from ideality. For example, the fugacity fugacity (α = 1 in Eq. 13), at least up to moderate pressures, and
coefficient of steam is 0.91 at 1073 K and 500 bars (Helgeson and indeed this has been experimentally documented (Dixon, et al.,
Kirkham, 1974) and that of CO2 is 1.2 at 1500 K and 500 bars 1995).
(Mel’nik, 1972), taking the standard state to be 1 bar. Higher Carbon solubility increases with increasing alkali and alkali
temperatures serve only to force fugacity coefficients closer to Earth metal contents in magmas, which is consistent with the
unity. CO2-H2O mixtures do have excess energies of mixing, but stability of carbonate mineral-like complexes in magmas (Lesne,
again, this effect is generally small at the high temperatures and et al., 2011; Duncan, et al., 2017). Higher temperatures favor the
moderate pressures of volatile-dominated atmospheres overlying dissolution of CO2 in the melt as carbonate (Konschak and Keppler,
magma oceans. The solution may require iteration between the 2014). Relatively large amounts of CO2 (molecular) can also dissolve

Frontiers in Earth Science 12 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 8
Distributions of H between atmosphere and magma ocean as functions of fO2 (ΔIW). Low-H (H equal to a 1 bar atmosphere of H2 or H2O) and high-
H scenarios (H equal to a 10 bar atmosphere of H2 or H2O) are considered. Masses of H in the low-H and high-H scenarios are broadly similar to those
inferred for the bulk silicate reservoirs of terrestrial bodies within the solar system. All scenarios assume an atmosphere of 100 bars total pressure, a
temperature of 2773 K, an Earth-sized planet, and a full mantle magma ocean. (A) Predictions of partial pressures of H2 and H2O gases in equilibrium
with a magma ocean. PH2O is nearly constant across the range of fO2 considered, whereas PH2 increases under more reducing conditions. (B)
Predictions of H species concentrations for magma oceans in equilibrium with partial pressures of gases in (A). (C) Mass ratios of H in the atmosphere
versus H dissolved in magma oceans. Oxidizing conditions force H2O stability and nearly all H to dissolve into magma oceans. Reducing conditions enable
a more equal distribution of H between atmosphere and magma ocean.

into more evolved (higher SiO2) melt compositions, following


Reaction 25. Higher temperatures also favor reactants in Reaction
25 (Konschak and Keppler, 2014), and it is therefore possible that
molecular CO2 contributes to CO2 solubility within magma oceans.
Sufficient experimental data have been collected to permit the
development of models for CO2 solubility in magmas (Dixon and
Stolper, 1995; Papale, et al., 2006; Iacono-Marziano, et al., 2012). We
again apply MAGMASAT to predict CO2 solubility in peridotitic
liquids over magma ocean P-T conditions (Gualda and Ghiorso,
2015) (Figure 9). Solubility scales with PCO2, as suggested by
Reactions 25 and 26 MAGMASAT predicts that higher
temperatures are predicted to decrease solubility, but the effect is
relatively minor, with solubility decreasing by a factor of 2 between
1,500 and 3000 K. Experimental work on mafic systems also
supports a relatively small effect of temperature; although there is
evidence that CO2 solubility increases with increasing temperature,
this is opposite to the predictions of MAGMASAT. Least squares
fitting of MAGMASAT predictions of CO2 solubility at 2273 K for 1,
10, 50, and 100 bars yields the following relationship for a BSE
FIGURE 9
magma composition: MAGMASAT predictions of CO2 solubility in an ultramafic melt as
CO2wt.%  2.14 × 10−4 · PCO2 (27) a function of PCO2 at different temperatures. The line is fit to the
2273 K data and is given by Eq. 27.
Under more reducing conditions, f CO2 will decrease while f CO
increases for any given temperature and pressure. With decreasing
f CO2, the carbonate and CO2 contents of a magma will without Fe, reacted with variable f CO have not found additional
correspondingly decrease, while C species associated with the evidence that Fe-carbonyl is a significant component of C solubility.
dissolution of CO will increase. Rather, there is building evidence that single CO groups are the
The exact nature of C species dissolved in magmas in dominant species (Armstrong, et al., 2015; Yoshioka, et al., 2019).
equilibrium with CO remains relatively uncertain. Higher than These groups may be molecular CO or CO species complexed to
expected C solubilities were observed in reduced, lunar analog other components in the melt. Without a more detailed
melts rich in Fe, leading to the suggestion that Fe-carbonyl understanding for how CO interacts with magma, extrapolation
groups could be a major component of CO solubility (Wetzel, of CO solubility determinations to magma ocean conditions will
et al., 2013). Other spectroscopic analyses of melts, with or introduce some uncertainty. Nonetheless, it is clear that C becomes

Frontiers in Earth Science 13 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

less soluble in magma as f CO2 falls and f CO rises. Yoshioka et al. (and f CO2) and reducing conditions promote graphite or diamond
(2019) developed the following equation to predict the solubility of saturation following the general reaction:
C (wt. %) in MORB magmas for a given f CO:
COg  0.5O2 (melt) + Ccrystal (31)
log(Cwt.% ) −5.20 + 0.80logf CO (28)
Once stabilized, a buffering phase such as diamond or graphite is
The unit for f CO is bar. Note that the solubility of CO is a new volatile reservoir that must be accounted for. The physical
expressed as reduced carbon, not CO, so statements of mass balance stability of the buffering phase must also be considered. For example,
that incorporate Eq. 28 must take this into account. graphite is expected to be buoyant in magma ocean liquids and could
The distribution of volatile carbon species between atmospheres therefore rise to the surface (Keppler and Golabek, 2019), whereas
and magma oceans are not predicted to be strongly dependent on volatile elements dissolved in alloys or stabilized in intermetallic
the amount of carbon present in the combined system as the compounds (carbide, nitrides, sulfides) may be dense and could
solubilities of all relevant species scale with their respective therefore sink within a magma ocean to ultimately join the core.
fugacities, with exponents close to unity.
Further decreases in f O2 will progressively destabilize CO, 3.1.3 Nitrogen
possibly resulting in graphite/diamond/carbide saturation if N2 is the dominant N-bearing species in the gas phase above
overall C contents are sufficiently high. Decreases in f O2 also magma oceans (Boulliung, et al., 2020) independent of fO2 (Dalou,
potentially help to stabilize methane or other CH-species if a et al., 2022; Sossi, et al., 2023). However, depending on fO2, the
source of H is also present (Figure 6): speciation of N in silicate melts simulating magma ocean conditions
falls into two domains. Under oxidized conditions, N dissolves as
2H2 g + COg  CH4 g + 0.5O2 g (29)
molecular N2 with a valence state of 0 and its solubility is primarily
and CH4 can then dissolve into the magma following: dependent on pressure following Henry’s Law (Libourel, et al., 2003;
Dalou, et al., 2019; Grewal, et al., 2020; Bernadou, et al., 2021). At
CH4 g  CH4 (melt) (30)
reduced conditions, N dissolves as ions with a −3 valence state
Once dissolved in a magma, methane may undergo further (Libourel et al., 2003; Dalou et al., 2019; Mosenfelder et al., 2019;
reactions to produce other hydrocarbon species. Spectroscopic Boulliung et al., 2020; Grewal et al., 2020; Bernadou et al., 2021).
observations identify CH4, or possibly methyl groups, dissolved Depending on fH2, these ions could either be anhydrous nitrides
in reduced, H-bearing silicate melts, supporting the importance of (N3−) or hydrous amines (NH2−) and ammonia (NH3) molecules
Reaction 29 (Mysen and Yamashita, 2010). Methane presumably (Mosenfelder, et al., 2019; Grewal, et al., 2020). N2 and NH3
dissolves into the ionic porosity of magma, following the example of molecules physically dissolve into the ionic porosity of silicate
H2 above, but how melt-reactive hydrocarbon species interact with melts whereas ionic nitrides and amides chemically dissolve into
the silicate network has not been well explored, and their importance the silicate melt network by displacing O2− from the silicate melt
to magma oceans is therefore not as well established as it is for other network (Libourel et al., 2003). Whereas N2, NH3, and NH2− species
species containing C or H. have been observed by Raman and FTIR (fourier transform
Figure 10 plots the distribution of C species between infrared) spectroscopy in quenched silicate glasses (Dalou, et al.,
atmospheres and magma oceans as a function of f O2 and the 2019; Mosenfelder, et al., 2019; Grewal, et al., 2020) (Figure 6), the
total abundance of C. Only solubility related to CO2 and CO is presence of anhydrous N3− has only been observed by Raman by
considered. We again assume a temperature of 2273 K, but total Dalou et al. (2022) at very low fO2. N solubility as N2 can be
pressure is now determined by the combined pressure of CO and represented as:
CO2. Atmospheric pressure therefore decreases as CO is converted N2 g  N2 (melt) (32)
to CO2 with increasing f O2. We explore a high-C scenario (mass
equivalent of 500 bars of C species in the atmosphere-magma ocean Even though hydrous N-H species are widely observed in
system) and a low-C scenario (mass equivalent 50 bars of C species experimental silicate melts simulating magma ocean conditions,
in the atmosphere-magma ocean system). For context, present-day the dissolution of N solely as N3− can adequately fit the observed
Venus has an atmosphere that is nearly 100 bars of CO2. The N abundances in reduced silicate melts in both hydrous and
crossover for partial pressures of CO and CO2 is near IW, but anhydrous conditions, and can be represented as:
higher pressures stabilize CO2 gas (Eq. 28) and shift the crossover to
3
more reducing conditions (Figure 10A). The crossover for the melt N2 g + 3O2− (melt)  2N3− (melt) + N2 g (33)
2
concentration of C related to CO and CO2 dissolution occurs below
IW because CO2 species are more soluble than those related to CO The solubility of N as molecular N2 under oxidized conditions
(Figure 10B). Magma oceans are the larger C reservoir when CO2 (Eq. 31) scales with fN2 and is independent of fO2, whereas the
dominates solubility, whereas the atmosphere is the larger reservoir solubility of N as N3− under reduced conditions (Eq. 32), in addition
when CO dominates solubility (Figure 10C). The ability of the to scaling with (fN2) with an exponent between 0.5 and 0.75.
atmosphere to dominate the C budget therefore appears Although the exact fO2 for the transition of N speciation
restricted to conditions more reducing than IW. depends upon several thermodynamic parameters like T, P, fluid
The discussion above for carbon assumes that the magma ocean composition, and melt composition, available evidence suggests that
remains undersaturated with respect to solid forms of C. High f CO this transition likely occurs near the IW redox buffer for silicate

Frontiers in Earth Science 14 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 10
Distributions of C between atmosphere and magma ocean as functions of fO2 (ΔIW). Low-C (C equal to a 50 bar atmosphere of CO or CO2) and
high-C scenarios (C equal to a 500 bar atmosphere of CO or CO2) are considered. Masses of C in the low-C and high-C scenarios are broadly similar to
those inferred for the bulk silicate reservoirs of terrestrial bodies within the solar system. All scenarios assume a temperature of 2773 K, an Earth-sized
planet, and a full mantle magma ocean. Atmospheric pressure varies with the total pressure of C-bearing gas molecules. (A) Predictions of partial
pressures of CO and CO2 gases in equilibrium with a magma ocean. PCO2 dominates above IW, whereas PCO dominates below IW. (B) Predictions of C
species concentrations for magma oceans in equilibrium with partial pressures of gases in (A). (C) Mass ratios of C in the atmosphere versus C dissolved in
magma oceans. Under oxidizing conditions, C is nearly equally distributed, but reducing conditions make the atmosphere the dominant C reservoir.

melts applicable to shallow magma ocean conditions (Boulliung 3.1.4 Sulfur


et al., 2020; Bernadou et al., 2021). It has long been known (Fincham and Richardson, 1954) that at
The solubility of N in analog magma ocean silicate melts has fO2 below the quartz-fayalite-magnetite buffer (<IW +3), S dissolves
been extensively studied by experimental studies over the last two as S2− by displacing O2− from the anion sublattice. Therefore, at
decades for a wide range of fO2, P, and T conditions. Libourel et al. oxygen fugacities relevant to magma oceans (generally <IW), S
(2003) calibrated the two-species model (dissolution of N as N2 and solubility in the silicate melts can be represented as:
N3− in oxidized and reduced conditions, respectively) with 1 1
experimental data at 1 bar to determine N solubility in basaltic O2− (melt) + S2 g  S2− (melt) + O2 g (35)
2 2
silicate melts as a function of PN and fO2:
The equilibrium of this equation can be described as:
Nppm  0.06PN2 + 5.97PN0.5 −0.75
2 f O2 (34)
lnKeq  lnaS2−  + lnfO0.5
2  − lnfS2  − lnaO 
0.5 2−
(36)
The first term on the righthand side accounts for N2 solubility,
The concentration of O2− anions in the silicate melt greatly
and the second term accounts for N3− solubility. Bernadou, et al.
exceeds those of other anions, including S2−. O2− concentration or
(2021) showed that the formalism for N solubility in the silicate melt
activity is assumed to be constant and Eq. 36 can be modified as
represented by Eq. 33 remains almost unaltered from 1 to 3,000 bars
(assuming the activity coefficient of S2− to be 1):
in the C-H-O-N system.
Figure 11 plots the distribution of N as a function of fO2 1 S2
ln(S) (melt)  ln (CS ) + lnf  (37)
between atmospheres and magma oceans. We explore a high-N 2 fO2
scenario (mass equivalent of 10 bars of N species in the
atmosphere-magma ocean system) and a low-N scenario (mass where CS , the sulfide capacity of the silicate melt, is a
equivalent of 1 bar of N species in the atmosphere-magma ocean pseudoequilbrium constant which is controlled by the
system). Atmospheres start to lose significant N while magma composition of the silicate melt, and primarily by its FeO content
oceans start to gain significant N below IW (Figure 11A and 9 (O’Neill and Mavrogenes, 2002). Using data from previous
(center)). Figure 11B shows that there are essentially two regimes experiments carried out at 1 atm in which a gas of known fO2
for N: i) above IW, nearly all N resides in atmospheres, as N2 is (between IW −1 and IW +3) and fS2 was equilibrated with silicate
relatively insoluble in melts; and ii) below IW −3, nearly all N melts, Gaillard et al. (2022) devised an empirical relationship to
resides in magma oceans, demonstrating the highly soluble determine S solubility in the silicate melts:
nature of nitride complexes. Between these two regimes is a 26476 1 S2
relatively narrow transition zone, although the fO2 values within lnSppm  13.84− +0.12FeOwt. % (melt) + lnf  (38)
T 2 fO2
this transitional zones overlap with those inferred for magma
oceans from studies of core formation. Atmospheres are the It should be noted that this equation is calibrated for FeO-rich
largest N reservoir when N2 controls solubility to ~IW −2, but terrestrial basalts with relatively simplified silicate melt
magma oceans quickly dominate the N budget as fO2 drops. compositions. Although new empirical models have been

Frontiers in Earth Science 15 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 11
Distributions of N between atmosphere and magma ocean as functions of fO2 (ΔIW). Low-N (N equal to a 1 bar atmosphere of N2) and high-N
scenarios (N equal to a 10 bar atmosphere of N2) are considered. Masses of N in the low-N and high-N scenarios are broadly similar to those inferred for
the bulk silicate reservoirs of terrestrial bodies within the solar system. All scenarios assume a temperature of 2773 K, an Earth-sized planet, and a full
mantle magma ocean. (A) Predictions of partial pressures of N2 gas in equilibrium with a magma ocean. (B) Predictions of N species concentrations
for magma oceans in equilibrium with partial pressures of N2 gas in (C). (C) Mass ratios of N in the atmosphere versus N dissolved in magma oceans. Under
oxidizing conditions, N nearly all N is present in the atmosphere, whereas reducing conditions force nearly all N to dissolve into magma oceans.

calibrated with experimental data for FeO-poor magmas (Namur, masses for the BSE in Figure 12B. Sulfur is highly soluble in magmas,
et al., 2016), their utility for atmosphere-magma ocean equilibration and this manifests as the much larger masses of S dissolved in
models is limited due the lack of a fS2 term. magma oceans compared to the equilibrium masses of S present in
The empirical equation has FeO and fO2 terms whose effects atmospheres (Figure 12A).
are counteract each other in Fe metal-saturated systems. Because Eq. 37 assumes that HS−species do not contribute to S solubility,
magma ocean silicate melts are in equilibrium with metal during as the compositional dependence of S solubility does not include a
core-mantle differentiation, their FeO contents are thought to be term related to fH2. With this assumption, the stabilization of H2S
lower than those of terrestrial basaltic magmas. Therefore, below gas with increasing fH2 under reducing conditions acts to make S
IW, the effect of the fO 2 term dominates such that the S more volatile. An example of this is provided in Figure 12 as H
solubility in the silicate melt increases at increasingly reduced scenario. There is debate as to whether HS−species are significant for
conditions. S solubility (Baker and Moretti, 2011), so Eq. 37 may underestimate
Sulfur solubility is typically expressed as a function of fS2 (Eq. S solubility in H2-rich systems. Nevertheless, it is clear that S is
37), but S2 is not necessarily the dominant S-bearing gas species over relatively soluble in melts and that primordial atmospheres contain
the P-T-X conditions of magma ocean-atmosphere exchange. only a small amount of S.
Indeed, S is unique in its tendency to form gas molecules with
other volatile elements. Hydrogen sulfide is produced by S2 when
exposed to elevated fH2, and COS is produced by S2 when exposed to 3.2 Beyond an equilibrium model
elevated fCO. This dynamic makes predicting the atmosphere-
magma ocean distribution of S correspondingly more The analysis above reviewed solubility laws and mass balance
complicated than for H, C, or N. statements to calculate the equilibrium model for the
Figure 12 plots the stability of S-bearing species for an distribution of H, C, N, and S between magma oceans and
atmosphere with 100 bars of total pressure. The gas phase is atmospheres. But this framework is not complete and much
10 mol% H, 0.03 mol% S, and the balance is C species. work is needed to make new experimental measurements of
Temperature is fixed at 2273 K, fO2 ranges between ΔIW −4 and solubility under the P-T-X conditions of magma oceans,
ΔIW +2. Hydrogen sulfide is the most abundant S-bearing gas, including high temperatures, high pressures, ultramafic melt
followed by COS, under reducing conditions, whereas SO2 is most compositions, and highly reducing conditions. Nevertheless,
abundant under oxidizing conditions. S2 remains a minority species applying an equilibrium model with a single P-T-X condition
over the entire range of fO2. for chemical exchange within a system as large and dynamic as a
Figure 12B plots the mass of S in magma oceans for our nominal combined magma ocean and atmosphere must fall short because
scenario (10 mol% H, 0.03 mol% S, and the balance is C species), a P-T-X conditions vary with time and heliocentric location as
high-S scenario (10 mol% H, 0.3 mol% S, and the balance C is planets grow.
species), and a high-H scenario (30 mol% H, 0.03 mol% S, and the As magma oceans cool, the interface temperature with the
balance C species). The mass of S in the magma ocean is calculated atmosphere will change (Lichtenberg, et al., 2021a; Lichtenberg,
using Eq. 37. All atmospheres are 100 bars total, and equilibrium is et al., 2022). The ability of a magma ocean and atmosphere to
calculated for 2273 K. The horizontal solid lines bracket estimated S remain in equilibrium relies on mass exchange across their

Frontiers in Earth Science 16 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

FIGURE 12
Distributions of sulfur between atmosphere and magma ocean as functions of fO2 (ΔIW). (A) An example calculation of S gas species partial pressures
for a 100-bar atmosphere with a nominal composition of 10 mol% H, 0.03 mol% S, and the balance is C species at 2273 K. (B) Predicted concentrations of
S (dissolved in magma oceans for a nominal atmosphere, a high-S atmosphere (10 mol% H, 0.3 mol% S, and the balance is C), and a high-H atmosphere
(30 mol% H, 0.03 mol% S, and the balance is C species). The horizontal lines bracket the estimated range of S concentrations in the bulk silicate Earth
(McDonough and Sun, 1995) (C) Mass ratios of S in the atmosphere versus S dissolved in magma oceans. All scenarios predict that most S is present within
magma oceans.

interface, but also the nature and associated kinetics of this constraints on the redox diversity and crystallization pathways of
exchange remain areas of study (Salvador, et al., 2017). The rocky planets within and beyond the Solar System.
diffusion of volatiles from the interface may be important here,
in addition to the convective patterns of the underlying magma
ocean and its associated timescales and length scales for mixing 4 Summary and outlook
(Solomatov, 2007). With continued cooling, magma oceans
crystallize, leading to a new era of evolution in which volatiles The cores of rocky planets could host significant amounts of
will partition between minerals and melts, depending on the volatile elements. However, a major finding from the recent
elemental exchange between the magma and its overlying metal-silicate partitioning literature is that high pressures and
atmosphere, and hence the formation of a protocrust (Bower, temperatures can have opposing effects on the partition
et al., 2022). Some minerals may be major hosts for volatiles, coefficients, and on average can lower the siderophile
but many likely exclude volatiles from their structures. tendencies of some elements including volatiles S and C. This
Crystallization should therefore lead to net outgassing. could result in decreased incorporation of these elements into Fe
Outgassing may be facilitated by bubble nucleation, bubble cores during late-stage core formation in larger rocky planets.
growth, and the diffusion of dissolved volatiles in the magma Though partial equilibration of large impactors during the giant
towards bubbles (Ikoma, et al., 2018). Again, kinetics may be impact phase of planetary growth could result in the transfer of
important in determining the chemical distribution of volatiles unequilibrated volatile-rich materials to cores. More
within the magma ocean-atmosphere system. Crystallization also measurements at high P-T are needed to ascertain trends for
offers the possibility of trapping buoyant magmas in the cumulate N and H. Both N and S appear to show increased affinity for iron
piles of magma oceans. Trapped magma provides a physical, rather while C partitions less into the metal at more oxidizing
than a chemical, mechanism for retaining volatiles in magma conditions. These change in siderophile behaviors of the
oceans (Hier-Majumder and Hirschmann, 2017) and is highly volatile elements has significant implications for the volatile
relevant for inhomogeneous crystallization paths in Earth- or reservoirs as planets grow. When coupled with accretion
super-Earth-sized exoplanets (Moore and Cowan, 2020). The models, the metal-silicate partitioning measurements support
amount of trapped liquid is also an expression of kinetic heterogeneous accretion scenarios whereby volatiles were
processes and the competition of melt migration timescales delivered towards the later stages of core formation. This is in
against crystallization timescales. Finally, recent work suggests contrast with the increasing evidence for the volatile-rich nature
that large impacts may result in transient, highly energetic of inner Solar System planetesimals which could have strongly
periods in which the distinction between atmosphere and influenced the volatile budgets of terrestrial planets. Though most
surface may vanish (Caracas and Stewart, 2023). Upon cooling, of the studies reviewed here were applied to the formation of the
these atmospheres would shift from being silicate-rich to rich in Earth and other rocky Solar System bodies, it has been inferred
moderately volatile elements. Ultimately, all of these processes that core segregation likely occur during the formation of rocky
must be accounted for to understand how much nature deviates planets in extrasolar planetary systems and in planets larger than
from the equilibrium framework. In light of associated degeneracy, the Earth such as super-earths and sub-neptunes (E.g., Otegi,
information from exoplanetary systems may provide additional et al., 2020). Calculations have indicated efficient mixing between

Frontiers in Earth Science 17 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

iron and rocky materials (Wahl and Militzer, 2015) and, similarly, Author contributions
volatiles and rocky materials (Dorn and Lichtenberg, 2021;
Kovačević, et al., 2022; Vazan, et al., 2022) would occur in T-AS, CJ, DG, and CD contributed to the collection and curation
such planets, and it would be beneficial to obtain of the data. All authors contributed to the article and approved the
measurements at higher P-T and a wider range of submitted version.
compositions than the current dataset to better understand
metal-rock and volatile-rock differentiation mechanisms in
super-Earths. Acknowledgments
In the magma ocean stage of growth, significant amounts of
volatiles can dissolve into the silicate parts of planets, in which The results reported herein benefitted from collaborations and/or
solubilities and redox state determine the availability of volatiles to information exchange within NASA’s Nexus for Exoplanet System
outgassing. The recent solubility studies indicate that under oxidizing Science (NExSS) research coordination network sponsored by NASA’s
conditions, most H and S are sequestered in the magma ocean, most Science Mission Directorate and project “Alien Earths” funded under
N is outgassed to the atmosphere, and C is nearly equally distributed Agreement No. 80NSSC21K0593. CRMJ acknowledges support from
between the atmosphere and the interior. Under reducing conditions, NASA Emerging Worlds Grant 80NSSC21K0377. TL was supported
nearly all N dissolves in the magma ocean, the atmosphere becomes by the Branco Weiss Foundation. CD acknowledges ANR grant JCJC
the dominant C reservoir, H becomes more equally distributed CSI Planet. DG was funded by a Barr Foundation Postdoctoral
between the interior and the atmosphere, and S remains Fellowship by California Institute of Technology. This work was
dominantly in the interior. Coupling of measurements with supported by the AEThER project, funded by the Alfred P. Sloan
microphysical to large-scale models is a direction for future work Foundation under grant No. G202114194. This material is based upon
that could provide further clarity on some of these multi-scale work supported by the Center for Matter at Atomic Pressures
processes. Finally, the understanding gained from these studies can (CMAP), a National Science Foundation (NSF) Physics Frontiers
guide the interpretation of new measurements of exoplanetary Center, under Award PHY2020249. The authors thank Vincent
atmospheres, which can in turn provide additional constraints to Clesi for sharing data tables for hydrogen.
assess the bulk compositions, formation histories, and the abilities of
rocky planets to develop conditions suitable for the origin and
sustainability of life as we know it. Conflict of interest
Recent first principles/ab initio molecular dynamics and
density functional theory (DFT) simulations have been used to The authors declare that the research was conducted in the
gain further insights on how volatile elements are distributed absence of any commercial or financial relationships that could be
among cores, magmas, and atmospheres. For example construed as a potential conflict of interest.
(Solomatova and Caracas, 2019; Davis, et al., 2022), found C The handling editor MR declared a past collaboration with the
coordination in silicate melts to be largely pressure dependent, author CD.
implying that terrestrial magma oceans could have contained
oxidized carbon polymers. The latter study also found strong
clustering between C and Fe, which increases as a function of Publisher’s note
pressure, suggesting C could be present in cores. Molecular
dynamics studies of Fe-N compounds lead to the conclusion All claims expressed in this article are solely those of the authors
that Earth’s core likely contains very low amounts of nitrogen and do not necessarily represent those of their affiliated organizations,
(Bajgain, et al., 2019), though this does not rule out higher N or those of the publisher, the editors and the reviewers. Any product
contents in the cores of other planets. DFT studies have been used that may be evaluated in this article, or claim that may be made by its
to argue for the preferential incorporation of H in the core (Li, manufacturer, is not guaranteed or endorsed by the publisher.
et al., 2020; Yuan and Steinle-Neumann, 2020). If validated by
meaurements, ab initio methods could provide a way to estimate
partition coefficients and solubilities at a wider range of Supplementary material
conditions than are achievable with current experiments. A
Bayesian inference approach might also provide a way to The Supplementary Material for this article can be found online
extrapolate these quantities to conditions beyond the at: https://www.frontiersin.org/articles/10.3389/feart.2023.1159412/
measurements (e.g., Gaffney, et al., 2022). full#supplementary-material

References
Albarede, F. (2009). Volatile accretion history of the terrestrial planets and dynamic Alexander, C. M. D. (2017). The origin of inner Solar System water. Philosophical
implications. Nature 461 (7268), 1227–1233. doi:10.1038/nature08477 Trans. R. Soc. A Math. Phys. Eng. Sci. 375 (2094), 20150384. doi:10.1098/rsta.2015.0384
Albarede, F., Ballhaus, C., Blichert-Toft, J., Lee, C. T., Marty, B., Moynier, F., et al. Alexander, C. M. D. (2019a). Quantitative models for the elemental and isotopic
(2013). Asteroidal impacts and the origin of terrestrial and lunar volatiles. Icarus 222 fractionations in chondrites: the carbonaceous chondrites. Geochimica Cosmochimica
(1), 44–52. doi:10.1016/j.icarus.2012.10.026 Acta 254, 277–309. doi:10.1016/j.gca.2019.02.008

Frontiers in Earth Science 18 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Alexander, C. M. D. (2019b). Quantitative models for the elemental and isotopic fugacity buffers. Earth Planet. Sci. Lett. 286 (3-4), 556–564. doi:10.1016/j.epsl.2009.
fractionations in the chondrites: the non-carbonaceous chondrites. Geochimica 07.022
Cosmochimica Acta 254, 246–276. doi:10.1016/j.gca.2019.01.026
Caracas, R., and Stewart, S. T. (2023). No magma ocean surface after giant impacts
Armstrong, K., Frost, D. J., McCammon, C. A., Rubie, D. C., and Boffa Ballaran, T. between rocky planets. Earth Planet. Sci. Lett. 608, 118014. doi:10.1016/j.epsl.2023.
(2019). Deep magma ocean formation set the oxidation state of Earth’s mantle. Science 118014
365 (6456), 903–906. doi:10.1126/science.aax8376
Carroll, M. R., and Stolper, E. M. (1993). Noble gas solubilities in silicate melts and
Armstrong, L. S., Hirschmann, M. M., Stanley, B. D., Falksen, E. G., and glasses: new experimental results for argon and the relationship between solubility and
Jacobsen, S. D. (2015). Speciation and solubility of reduced C–O–H–N volatiles ionic porosity. Geochimica Cosmochimica Acta 57 (23-24), 5039–5051. doi:10.1016/
in mafic melt: implications for volcanism, atmospheric evolution, and deep volatile 0016-7037(93)90606-w
cycles in the terrestrial planets. Geochimica Cosmochimica Acta 171, 283–302.
Catling, D. C., Zahnle, K. J., and McKay, C. P. (2001). Biogenic methane, hydrogen
doi:10.1016/j.gca.2015.07.007
escape, and the irreversible oxidation of early Earth. Science 293 (5531), 839–843. doi:10.
Badding, J. V., Hemley, R. J., and Mao, H. K. (1991). High-pressure chemistry of 1126/science.1061976
hydrogen in metals: in situ study of iron hydride. Science 253 (5018), 421–424. doi:10.
Chabot, N. L. (2004). Sulfur contents of the parental metallic cores of magmatic iron
1126/science.253.5018.421
meteorites. Geochimica Cosmochimica Acta 68 (17), 3607–3618. doi:10.1016/j.gca.2004.
Badro, J., Alexander, S. C., and John, P. B. (2014). A seismologically consistent 03.023
compositional model of Earth’s core. Proc. Natl. Acad. Sci. 111 (21), 7542–7545. doi:10.
Chabot, N. L., and Agee, C. B. (2003). Core formation in the Earth and Moon: new
1073/pnas.1316708111
experimental constraints from V, Cr, and Mn. Geochimica Cosmochimica Acta 67 (11),
Bajgain, S. K., Mookherjee, M., Dasgupta, R., Ghosh, D. B., and Karki, B. B. (2019). 2077–2091. doi:10.1016/s0016-7037(02)01272-3
Nitrogen content in the Earth’s outer core. Geophys. Res. Lett. 46 (1), 89–98. doi:10.
Chachan, Y., and Stevenson, D. J. (2018). On the role of dissolved gases in the
1029/2018gl080555
atmosphere retention of low-mass low-density planets. Astrophysical J. 854 (1), 21.
Baker, D. R., and Moretti, R. (2011). Modeling the solubility of sulfur in magmas: a 50- doi:10.3847/1538-4357/aaa459
year old geochemical challenge. Rev. Mineralogy Geochem. 73 (1), 167–213. doi:10.2138/
Chambers, J. E., and Wetherill, G. W. (1998). Making the terrestrial planets: N-body
rmg.2011.73.7
integrations of planetary embryos in three dimensions. Icarus 136 (2), 304–327. doi:10.
Bar-Nun, A., and Owen, T. (1998). Trapping of gases in water ice and consequences to 1006/icar.1998.6007
comets and the atmospheres of the inner planets. Sol. Syst. Ices 1998, 353–366.
Chao, K.-H., deGraffenried, R., Lach, M., Nelson, W., Truax, K., and Gaidos, E.
Batalha, N. M. (2014). Exploring exoplanet populations with NASA’s Kepler Mission. (2021). Lava worlds: from early earth to exoplanets. Geochemistry 81 (2), 125735. doi:10.
Proc. Natl. Acad. Sci. 111 (35), 12647–12654. doi:10.1073/pnas.1304196111 1016/j.chemer.2020.125735
Bernadou, F., Gaillard, F., Füri, E., Marrocchi, Y., and Slodczyk, A. (2021). Nitrogen Chi, H., Dasgupta, R., Duncan, M. S., and Shimizu, N. (2014). Partitioning of carbon
solubility in basaltic silicate melt - implications for degassing processes. Chem. Geol. between Fe-rich alloy melt and silicate melt in a magma ocean–implications for the
573, 120192. doi:10.1016/j.chemgeo.2021.120192 abundance and origin of volatiles in Earth, Mars, and the Moon. Geochimica
Cosmochimica Acta 139, 447–471. doi:10.1016/j.gca.2014.04.046
Blanchard, I., Rubie, D., Jennings, E., Franchi, I., Zhao, X., Petitgirard, S., et al. (2022).
The metal–silicate partitioning of carbon during Earth’s accretion and its distribution in Chidester, B. A., Lock, S. J., Swadba, K. E., Rahman, Z., Righter, K., and Campbell, A. J.
the early solar system. Earth Planet. Sci. Lett. 580, 117374. doi:10.1016/j.epsl.2022. (2022). The lithophile element budget of Earth’s core. Geochem. Geophys. Geosystems 23
117374 (2), e2021GC009986. doi:10.1029/2021gc009986
Blanchard, I., Siebert, J., Borensztajn, S., and Badro, J. (2017). The solubility of heat- Clesi, V., Bouhifd, M. A., Bolfan-Casanova, N., Manthilake, G., Schiavi, F., Raepsaet,
producing elements in Earth’s core. Geochem. Perspect. Lett. 5, 1–5. doi:10.7185/ C., et al. (2018). Low hydrogen contents in the cores of terrestrial planets. Sci. Adv. 4 (3),
geochemlet.1737 e1701876. doi:10.1126/sciadv.1701876
Bondar, D., Withers, A. C., Whittington, A. G., Fei, H., and Katsura, T. (2023). Corgne, A., Keshav, S., Wood, B. J., McDonough, W. F., and Fei, Y. (2008a).
Dissolution mechanisms of water in depolymerized silicate (peridotitic) glasses based Metal–silicate partitioning and constraints on core composition and oxygen fugacity
on infrared spectroscopy. Geochimica Cosmochimica Acta 342, 45–61. doi:10.1016/j.gca. during Earth accretion. Geochimica Cosmochimica Acta 72 (2), 574–589. doi:10.1016/j.
2022.11.029 gca.2007.10.006
Bonsor, A., Carter, P. J., Hollands, M., Gänsicke, B. T., Leinhardt, Z., and Harrison, Corgne, A., Siebert, J., and Badro, J. (2009). Oxygen as a light element: a solution to
J. H. D. (2020). Are exoplanetesimals differentiated? Mon. Notices R. Astronomical Soc. single-stage core formation. Earth Planet. Sci. Lett. 288 (1-2), 108–114. doi:10.1016/j.
492 (2), 2683–2697. doi:10.1093/mnras/stz3603 epsl.2009.09.012
Bonsor, A., Lichtenberg, T., Drazkowska, J., and Buchan, A. M. (2022). Rapid Corgne, A., Wood, B. J., and Fei, Y. (2008b). C-and S-rich molten alloy immiscibility
formation of exoplanetesimals revealed by white dwarfs. Nat. Astron. 7, 39–48. and core formation of planetesimals. Geochimica Cosmochimica Acta 72 (9),
doi:10.1038/s41550-022-01815-8 2409–2416. doi:10.1016/j.gca.2008.03.001
Bouhifd, M. A., and Jephcoat, A. P. (2011). Convergence of Ni and Co metal–silicate Dahl, T. W., and Stevenson, D. J. (2010). Turbulent mixing of metal and silicate
partition coefficients in the deep magma-ocean and coupled silicon–oxygen solubility in during planet accretion—and interpretation of the Hf–W chronometer. Earth Planet.
iron melts at high pressures. Earth Planet. Sci. Lett. 307 (3-4), 341–348. doi:10.1016/j. Sci. Lett. 295 (1-2), 177–186. doi:10.1016/j.epsl.2010.03.038
epsl.2011.05.006
Dalou, C., Deligny, C., and Füri, E. (2022). Nitrogen isotope fractionation during
Boujibar, A., Andrault, D., Bouhifd, M. A., Bolfan-Casanova, N., Devidal, J. L., and magma ocean degassing: tracing the composition of early Earth’s atmosphere. Geochem.
Trcera, N. (2014). Metal–silicate partitioning of sulphur, new experimental and Perspect. Lett. 20, 27–31. doi:10.7185/geochemlet.2204
thermodynamic constraints on planetary accretion. Earth Planet. Sci. Lett. 391,
Dalou, C., Füri, E., Deligny, C., Piani, L., Caumon, M. C., Laumonier, M., et al. (2019).
42–54. doi:10.1016/j.epsl.2014.01.021
Redox control on nitrogen isotope fractionation during planetary core formation. Proc.
Boujibar, A., Driscoll, P., and Fei, Y. (2020). Super-Earth internal structures and initial Natl. Acad. Sci. 116 (29), 14485–14494. doi:10.1073/pnas.1820719116
thermal states. J. Geophys. Res. Planets 125 (5), e2019JE006124. doi:10.1029/
Dalou, C., Hirschmann, M. M., von der Handt, A., Mosenfelder, J., and Armstrong, L.
2019je006124
S. (2017). Nitrogen and carbon fractionation during core–mantle differentiation at
Boulliung, J., Füri, E., Dalou, C., Tissandier, L., Zimmermann, L., and Marrocchi, Y. shallow depth. Earth Planet. Sci. Lett. 458, 141–151. doi:10.1016/j.epsl.2016.10.026
(2020). Oxygen fugacity and melt composition controls on nitrogen solubility in
Dasgupta, R., Chi, H., Shimizu, N., Buono, A. S., and Walker, D. (2013). Carbon
silicate melts. Geochimica Cosmochimica Acta 284, 120–133. doi:10.1016/j.gca.2020.
solution and partitioning between metallic and silicate melts in a shallow magma ocean:
06.020
implications for the origin and distribution of terrestrial carbon. Geochimica
Bower, D. J., Hakim, K., Sossi, P. A., and Sanan, P. (2022). Retention of water in Cosmochimica Acta 102, 191–212. doi:10.1016/j.gca.2012.10.011
terrestrial magma oceans and carbon-rich early atmospheres. Planet. Sci. J. 3 (4), 93.
Dasgupta, R., and Grewal, D. S. (2019). Origin and early differentiation of carbon and
doi:10.3847/psj/ac5fb1
associated life-essential volatile elements on Earth. Deep carbon 2019, 4–39. doi:10.
Brennan, M. C., Fischer, R. A., and Irving, J. C. E. (2020). Core formation and 1017/9781108677950.002
geophysical properties of Mars. Earth Planet. Sci. Lett. 530, 115923. doi:10.1016/j.epsl.
Dasgupta, R., and Hirschmann, M. M. (2010). The deep carbon cycle and melting in
2019.115923
Earth’s interior. Earth Planet. Sci. Lett. 298 (1-2), 1–13. doi:10.1016/j.epsl.2010.06.039
Busemann, H., Lorenzetti, S., and Eugster, O. (2006). Noble gases in D’Orbigny,
Davies, E. J., Carter, P. J., Root, S., Kraus, R. G., Spaulding, D. K., Stewart, S. T., et al.
Sahara 99555 and D’Orbigny glass—evidence for early planetary processing on the
(2020). Silicate melting and vaporization during rocky Planet Formation. J. Geophys.
angrite parent body. Geochimica cosmochimica acta 70 (21), 5403–5425. doi:10.1016/j.
Res. Planets 125 (2), e2019JE006227. doi:10.1029/2019je006227
gca.2006.08.015
Davis, A. H., Solomatova, N. V., Campbell, A. J., and Caracas, R. (2022). The
Campbell, A. J., Danielson, L., Righter, K., Seagle, C. T., Wang, Y., and Prakapenka, V.
speciation and coordination of a deep earth carbonate-silicate-metal melt.
B. (2009). High pressure effects on the iron–iron oxide and nickel–nickel oxide oxygen
J. Geophys. Res. Solid Earth 127 (3), e2021JB023314. doi:10.1029/2021jb023314

Frontiers in Earth Science 19 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Deguen, R., Landeau, M., and Olson, P. (2014). Turbulent metal–silicate mixing, Grewal, D. S., Dasgupta, R., and Farnell, A. (2020). The speciation of carbon, nitrogen,
fragmentation, and equilibration in magma oceans. Earth Planet. Sci. Lett. 391, 274–287. and water in magma oceans and its effect on volatile partitioning between major
doi:10.1016/j.epsl.2014.02.007 reservoirs of the Solar System rocky bodies. Geochimica Cosmochimica Acta 280,
281–301. doi:10.1016/j.gca.2020.04.023
Deligny, C., Füri, E., and Deloule, E. (2021). Origin and timing of volatile delivery (N,
H) to the angrite parent body: constraints from in situ analyses of melt inclusions. Grewal, D. S., Dasgupta, R., Holmes, A. K., Costin, G., Li, Y., and Tsuno, K. (2019a).
Geochimica Cosmochimica Acta 313, 243–256. doi:10.1016/j.gca.2021.07.038 The fate of nitrogen during core-mantle separation on Earth. Geochimica cosmochimica
acta 251, 87–115. doi:10.1016/j.gca.2019.02.009
Ding, F., and Wordsworth, R. D. (2022). Prospects for water vapor detection in the
atmospheres of temperate and arid rocky exoplanets around M-dwarf stars. Grewal, D. S., Dasgupta, R., Hough, T., and Farnell, A. (2021b). Rates of
Astrophysical J. Lett. 925 (1), L8. doi:10.3847/2041-8213/ac4a5d protoplanetary accretion and differentiation set nitrogen budget of rocky planets.
Nat. Geosci. 14 (6), 369–376. doi:10.1038/s41561-021-00733-0
Dixon, J. E., and Stolper, E. M. (1995). An experimental study of water and carbon
dioxide solubilities in mid-ocean ridge basaltic liquids. Part II: applications to degassing. Grewal, D. S., Dasgupta, R., and Marty, B. (2021c). A very early origin of isotopically
J. petrology 36 (6), 1633–1646. distinct nitrogen in inner Solar System protoplanets. Nat. Astron. 5 (4), 356–364. doi:10.
1038/s41550-020-01283-y
Dixon, J. E., Stolper, E. M., and Holloway, J. R. (1995). An experimental study of water
and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part I: calibration Grewal, D. S., Dasgupta, R., Sun, C., Tsuno, K., and Costin, G. (2019b). Delivery of
and solubility models. J. Petrology 36 (6), 1607–1631. carbon, nitrogen, and sulfur to the silicate Earth by a giant impact. Sci. Adv. 5 (1),
eaau3669. doi:10.1126/sciadv.aau3669
Dorn, C., and Lichtenberg, T. (2021). Hidden water in magma ocean exoplanets.
Astrophysical J. Lett. 922 (1), L4. doi:10.3847/2041-8213/ac33af Grewal, D. S., Seales, J. D., and Dasgupta, R. (2022a). Internal or external magma
oceans in the earliest protoplanets–Perspectives from nitrogen and carbon
Dreibus, G., and Wanke, H. (1985). Mars, a volatile-rich planet. Meteoritics 20,
fractionation. Earth Planet. Sci. Lett. 598, 117847. doi:10.1016/j.epsl.2022.117847
367–381.
Grewal, D. S., Sun, T., Aithala, S., Hough, T., Dasgupta, R., Yeung, L. Y., et al. (2022b).
Duncan, M. S., Dasgupta, R., and Tsuno, K. (2017). Experimental determination of
Limited nitrogen isotopic fractionation during core-mantle differentiation in rocky
CO2 content at graphite saturation along a natural basalt-peridotite melt join:
protoplanets and planets. Geochimica Cosmochimica Acta 338, 347–364. doi:10.1016/j.
implications for the fate of carbon in terrestrial magma oceans. Earth Planet. Sci.
gca.2022.10.025
Lett. 466, 115–128. doi:10.1016/j.epsl.2017.03.008
Gualda, G. A. R., and Ghiorso, M. S. (2015). MELTS _ E xcel: AM icrosoft E xcel-
Elkins-Tanton, L. T. (2012). Magma oceans in the inner solar system. Annu. Rev.
based MELTS interface for research and teaching of magma properties and
Earth Planet. Sci. 40, 113–139. doi:10.1146/annurev-earth-042711-105503
evolution. Geochem. Geophys. Geosystems 16 (1), 315–324. doi:10.1002/
Fei, Y., Prewitt, C. T., Mao, H. k., and Bertka, C. M. (1995). Structure and density of 2014gc005545
FeS at high pressure and high temperature and the internal structure of Mars. Science
Halliday, A. N. (2013). The origins of volatiles in the terrestrial planets. Geochimica
268 (5219), 1892–1894. doi:10.1126/science.268.5219.1892
Cosmochimica Acta 105, 146–171. doi:10.1016/j.gca.2012.11.015
Fichtner, C. E., Schmidt, M. W., Liebske, C., Bouvier, A. S., and Baumgartner, L. P.
Helgeson, H. C., and Kirkham, D. H. (1974). Theoretical prediction of the
(2021). Carbon partitioning between metal and silicate melts during Earth accretion.
thermodynamic behavior of aqueous electrolytes at high pressures and
Earth Planet. Sci. Lett. 554, 116659. doi:10.1016/j.epsl.2020.116659
temperatures; I, Summary of the thermodynamic/electrostatic properties of the
Fincham, C. J. B., and Richardson, F. D. (1954). The behaviour of sulphur in silicate solvent. Am. J. Sci. 274 (10), 1089–1198. doi:10.2475/ajs.274.10.1089
and aluminate melts. Proc. R. Soc. Lond. Ser. A. Math. Phys. Sci. 223 (1152), 40–62.
Hier-Majumder, S., and Hirschmann, M. M. (2017). The origin of volatiles in the E
Fine, G., and Stolper, E. (1986). Dissolved carbon dioxide in basaltic glasses: arth’s mantle. Geochem. Geophys. Geosystems 18 (8), 3078–3092. doi:10.1002/
concentrations and speciation. Earth Planet. Sci. Lett. 76 (3-4), 263–278. doi:10. 2017gc006937
1016/0012-821x(86)90078-6
Hirschmann, M. M. (2012). Magma ocean influence on early atmosphere mass and
Fischer, R. A., Cottrell, E., Hauri, E., Lee, K. K. M., and Le Voyer, M. (2020). The composition. Earth Planet. Sci. Lett. 341, 48–57. doi:10.1016/j.epsl.2012.06.015
carbon content of Earth and its core. Proc. Natl. Acad. Sci. 117 (16), 8743–8749. doi:10.
Hirschmann, M. M. (2016). Constraints on the early delivery and fractionation of
1073/pnas.1919930117
Earth’s major volatiles from C/H, C/N, and C/S ratios. Am. Mineralogist 101 (3),
Fischer, R. A., Nakajima, Y., Campbell, A. J., Frost, D. J., Harries, D., Langenhorst, F., 540–553. doi:10.2138/am-2016-5452
et al. (2015). High pressure metal–silicate partitioning of Ni, Co, V, Cr, Si, and O.
Hirschmann, M. M. (2018). Comparative deep Earth volatile cycles: the case for C
Geochimica Cosmochimica Acta 167, 177–194. doi:10.1016/j.gca.2015.06.026
recycling from exosphere/mantle fractionation of major (H2O, C, N) volatiles and from
Fulton, B. J., Petigura, E. A., Howard, A. W., Isaacson, H., Marcy, G. W., Cargile, P. A., H2O/Ce, CO2/Ba, and CO2/Nb exosphere ratios. Earth Planet. Sci. Lett. 502, 262–273.
et al. (2017). The California-Kepler survey. III. A gap in the radius distribution of small doi:10.1016/j.epsl.2018.08.023
planets. Astronomical J. 154 (3), 109. doi:10.3847/1538-3881/aa80eb
Huang, D., and Badro, J. (2018). Fe-Ni ideality during core formation on Earth. Am.
Gaffney, J. A., Yang, L., and Ali, S. (2022). Constraining model uncertainty in plasma Mineralogist 103 (10), 1707–1710. doi:10.2138/am-2018-6651
equation-of-state models with a physics-constrained Gaussian process. arXiv preprint
Iacono-Marziano, G., Morizet, Y., Le Trong, E., and Gaillard, F. (2012). New
arXiv:2207.00668. Available at: https://doi.org/10.48550/arXiv.2207.00668.
experimental data and semi-empirical parameterization of H2O–CO2 solubility in
Gaillard, F., Bernadou, F., Roskosz, M., Bouhifd, M. A., Marrocchi, Y., Iacono- mafic melts. Geochimica Cosmochimica Acta 97, 1–23. doi:10.1016/j.gca.2012.
Marziano, G., et al. (2022). Redox controls during magma ocean degassing. Earth 08.035
Planet. Sci. Lett. 577, 117255. doi:10.1016/j.epsl.2021.117255
Iacono-Marziano, G., Paonita, A., Rizzo, A., Scaillet, B., and Gaillard, F. (2010). Noble
Gaillard, F., Bouhifd, M. A., Füri, E., Malavergne, V., Marrocchi, Y., Noack, L., et al. gas solubilities in silicate melts: new experimental results and a comprehensive model of
(2021). The diverse planetary ingassing/outgassing paths produced over billions of years the effects of liquid composition, temperature and pressure. Chem. Geol. 279 (3-4),
of magmatic activity. Space Sci. Rev. 217 (1), 22–54. doi:10.1007/s11214-021-00802-1 145–157. doi:10.1016/j.chemgeo.2010.10.017
Ganguly, J. (2008). Thermodynamics in earth and planetary sciences. Heidelberg: Iacovino, K., Matthews, S., Wieser, P. E., Moore, G. M., and Bégué, F. (2021). VESIcal
Springer. Part I: an open-source thermodynamic model engine for mixed volatile (H2O-CO2)
solubility in silicate melts. Earth Space Sci. 8 (11), e2020EA001584. doi:10.1029/
Ghiorso, M. S., and Gualda, G. A. R. (2015). An H 2 O–CO 2 mixed fluid saturation
2020ea001584
model compatible with rhyolite-MELTS. Contributions Mineralogy Petrology 169,
53–30. doi:10.1007/s00410-015-1141-8 Ih, J., Kempton, E. M. R., Whittaker, E. A., and Lessard, M. (2023). Constraining the
thickness of TRAPPIST-1 b’s atmosphere from its JWST secondary eclipse observation
Gounelle, M., and Zolensky, M. E. (2014). The Orgueil meteorite: 150 years of history.
at 15 μm. Astrophysical J. Lett. 952 (1), L4. doi:10.3847/2041-8213/ace03b
Meteorit. Planet. Sci. 49 (10), 1769–1794. doi:10.1111/maps.12351
Ikoma, M., Elkins-Tanton, L., Hamano, K., and Suckale, J. (2018). Water partitioning
Greene, T. P., Bell, T. J., Ducrot, E., Dyrek, A., Lagage, P. O., and Fortney, J. J. (2023).
in planetary embryos and protoplanets with magma oceans. Space Sci. Rev. 214, 76–28.
Thermal emission from the Earth-sized exoplanet TRAPPIST-1 b using JWST. Nature
doi:10.1007/s11214-018-0508-3
618 (7963), 39–42. doi:10.1038/s41586-023-05951-7
Jackson, C. R., Bennett, N. R., Du, Z., Cottrell, E., and Fei, Y. (2018). Early episodes of
Grewal, D. S. (2022). Origin of nitrogen isotopic variations in the rocky bodies of the
high-pressure core formation preserved in plume mantle. Nature 553, 491.
solar system. Astrophysical J. 937 (2), 123. doi:10.3847/1538-4357/ac8eb4
Jackson, C. R. M., Cottrell, E., Du, Z., Bennett, N., and Fei, Y. (2021). High pressure
Grewal, D. S., and Asimow, P. D. (2023). Origin of the superchondritic carbon/
redistribution of nitrogen and sulfur during planetary stratification. Geochem. Perspect.
nitrogen ratio of the bulk silicate Earth− an outlook from iron meteorites. Geochimica
Lett. 18, 37–42. doi:10.7185/geochemlet.2122
Cosmochimica Acta 344, 146–159. doi:10.1016/j.gca.2023.01.012
Jambon, A., Weber, H., and Braun, O. (1986). Solubility of He, Ne, Ar, Kr and Xe in a
Grewal, D. S., Dasgupta, R., and Aithala, S. (2021a). The effect of carbon
basalt melt in the range 1250–1600 C. Geochemical implications. Geochimica
concentration on its core-mantle partitioning behavior in inner Solar System rocky
Cosmochimica Acta 50 (3), 401–408. doi:10.1016/0016-7037(86)90193-6
bodies. Earth Planet. Sci. Lett. 571, 117090. doi:10.1016/j.epsl.2021.117090

Frontiers in Earth Science 20 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Jana, D., and Walker, D. (1997). The influence of sulfur on partitioning of siderophile Lichtenberg, T. (2022). Redox hysteresis of super-Earth exoplanets from magma
elements. Geochimica Cosmochimica Acta 61 (24), 5255–5277. doi:10.1016/s0016- ocean circulation. Astrophysical J. Lett. 914 (1), L4. doi:10.3847/2041-8213/ac0146
7037(97)00307-4
Lichtenberg, T., Bower, D. J., Hammond, M., Boukrouche, R., Sanan, P., Tsai, S., et al.
Johansen, A., Ronnet, T., Bizzarro, M., Schiller, M., Lambrechts, M., (2021a). Vertically resolved magma ocean–protoatmosphere evolution: H2, H2O, CO2,
Nordlund, Å., et al. (2021). A pebble accretion model for the formation of CH4, CO, O2, and N2 as primary absorbers. J. Geophys. Res. Planets 126 (2),
the terrestrial planets in the Solar System. Sci. Adv. 7 (8), eabc0444. doi:10.1126/ e2020JE006711. doi:10.1029/2020je006711
sciadv.abc0444
Lichtenberg, T., Drażkowska, J., Schönbächler, M., Golabek, G. J., and Hands, T. O.
Jones, J. H., and Drake, M. J. (1986). Geochemical constraints on core formation in (2021b). Bifurcation of planetary building blocks during Solar System formation.
the Earth. Nature 322 (6076), 221–228. doi:10.1038/322221a0 Science 371 (6527), 365–370. doi:10.1126/science.abb3091
Kaltenegger, L., Pepper, J., Stassun, K., and Oelkers, R. (2019). TESS habitable zone Lichtenberg, T., Golabek, G. J., Burn, R., Meyer, M. R., Alibert, Y., Gerya, T. V., et al.
star catalog. Astrophysical J. Lett. 874 (1), L8. doi:10.3847/2041-8213/ab0e8d (2019). A water budget dichotomy of rocky protoplanets from 26Al-heating. Nat.
Astron. 3 (4), 307–313. doi:10.1038/s41550-018-0688-5
Karato, S.-I., and Murthy, V. R. (1997). Core formation and chemical equilibrium in
the Earth—I. Physical considerations. Phys. Earth Planet. Interiors 100 (1-4), 61–79. Lichtenberg, T., and Krijt, S. (2021). System-level fractionation of carbon from disk
doi:10.1016/s0031-9201(96)03232-3 and planetesimal processing. Astrophysical J. Lett. 913 (2), L20. doi:10.3847/2041-8213/
abfdce
Kempton, E. M. R., Lessard, M., Malik, M., Rogers, L. A., Futrowsky, K. E., Ih, J., et al.
(2023). Where are the water worlds? self-consistent models of water-rich exoplanet Lichtenberg, T., Schaefer, L. K., Nakajima, M., and Fischer, R. A. (2022). Geophysical
atmospheres. Astrophysical J. 953 (1), 57. doi:10.3847/1538-4357/ace10d evolution during rocky Planet Formation. arXiv preprint arXiv:2203.10023. Available
at: https://doi.org/10.48550/arXiv.2203.10023.
Keppler, H., and Golabek, G. (2019). Graphite floatation on a magma ocean and the
fate of carbon during core formation. Geochem. Perspect. Lett. 11, 12–17. doi:10.7185/ Luth, R. W., O Mysen, B., and Virgo, D. (1987). Raman spectroscopic study of the
geochemlet.1918 solubility behavior of H2 in the system Na2O-Al2O3-SiO2-H2. Am. Mineralogist 72 (5-
6), 481–486.
Kimura, T., and Ikoma, M. (2022). Predicted diversity in water content of terrestrial
exoplanets orbiting M dwarfs. Nat. Astron. 6 (11), 1296–1307. doi:10.1038/s41550-022- Lv, C., and Liu, J. (2022). Early planetary processes and light elements in iron-
01781-1 dominated cores. Acta Geochim. 41 (4), 625–649. doi:10.1007/s11631-021-00522-x
Kite, E. S., Fegley Jr., B., Schaefer, L., and Ford, E. B. (2020). Atmosphere origins for Mahan, B., Siebert, J., Blanchard, I., Badro, J., Kubik, E., Sossi, P., et al. (2018).
exoplanet sub-neptunes. Astrophysical J. 891 (2), 111. doi:10.3847/1538-4357/ab6ffb Investigating earth’s formation history through copper and sulfur metal-silicate
partitioning during core-mantle differentiation. J. Geophys. Res. Solid Earth 123
Kite, E. S., and Schaefer, L. (2021). Water on hot rocky exoplanets. Astrophysical
(10), 8349–8363. doi:10.1029/2018jb015991
J. Lett. 909 (2), L22. doi:10.3847/2041-8213/abe7dc
Malavergne, V., Bureau, H., Raepsaet, C., Gaillard, F., Poncet, M., Surblé, S., et al.
Konschak, A., and Keppler, H. (2014). The speciation of carbon dioxide in silicate
(2019). Experimental constraints on the fate of H and C during planetary core-mantle
melts. Contributions Mineralogy Petrology 167, 998–1013. doi:10.1007/s00410-014-
differentiation. Implications for the Earth. Icarus 321, 473–485. doi:10.1016/j.icarus.
0998-2
2018.11.027
Kovačević, T., González-Cataldo, F., Stewart, S. T., and Militzer, B. (2022). Miscibility
Mann, U., Frost, D. J., and Rubie, D. C. (2009). Evidence for high-pressure core-
of rock and ice in the interiors of water worlds. Sci. Rep. 12 (1), 13055. doi:10.1038/
mantle differentiation from the metal–silicate partitioning of lithophile and weakly-
s41598-022-16816-w
siderophile elements. Geochimica Cosmochimica Acta 73 (24), 7360–7386. doi:10.1016/
Kuwahara, H., Itoh, S., Nakada, R., and Irifune, T. (2019). The effects of carbon j.gca.2009.08.006
concentration and silicate composition on the metal-silicate partitioning of carbon in a
Mansfield, M., Kite, E. S., Hu, R., Koll, D. D. B., Malik, M., Bean, J. L., et al. (2019).
shallow magma ocean. Geophys. Res. Lett. 46 (16), 9422–9429. doi:10.1029/
Identifying atmospheres on rocky exoplanets through inferred high albedo.
2019gl084254
Astrophysical J. 886 (2), 141. doi:10.3847/1538-4357/ab4c90
Kuwahara, H., Itoh, S., Suzumura, A., Nakada, R., and Irifune, T. (2021). Nearly
Marty, B. (2012). The origins and concentrations of water, carbon, nitrogen and noble
carbon-saturated magma oceans in planetary embryos during core formation. Geophys.
gases on Earth. Earth Planet. Sci. Lett. 313, 56–66. doi:10.1016/j.epsl.2011.10.040
Res. Lett. 48 (10), e2021GL092389. doi:10.1029/2021gl092389
Matsui, T., and Abe, Y. (1986). Evolution of an impact-induced atmosphere and
Lesne, P., Kohn, S. C., Blundy, J., Witham, F., Botcharnikov, R. E., and Behrens, H.
magma ocean on the accreting Earth. Nature 319 (6051), 303–305. doi:10.1038/
(2011). Experimental simulation of closed-system degassing in the system
319303a0
basalt–H2O–CO2–S–Cl. J. Petrology 52 (9), 1737–1762. doi:10.1093/petrology/
egr027 Mavrogenes, J. A., and O’Neill, H.St C. (1999). The relative effects of pressure,
temperature and oxygen fugacity on the solubility of sulfide in mafic magmas.
Lewis, J. A., Jones, R. H., and Brearley, A. J. (2022). Plagioclase alteration and
Geochimica Cosmochimica Acta 63 (7-8), 1173–1180. doi:10.1016/s0016-7037(98)
equilibration in ordinary chondrites: metasomatism during thermal metamorphism.
00289-0
Geochimica Cosmochimica Acta 316, 201–229. doi:10.1016/j.gca.2021.10.004
McCubbin, F. M., and Barnes, J. J. (2019). Origin and abundances of H2O in the
Li, J., and Agee, C. B. (1996). Geochemistry of mantle–core differentiation at high
terrestrial planets, Moon, and asteroids. Earth Planet. Sci. Lett. 526, 115771. doi:10.
pressure. Nature 381 (6584), 686–689. doi:10.1038/381686a0
1016/j.epsl.2019.115771
Li, J., and Agee, C. B. (2001). Element partitioning constraints on the light element
McDonough, W. F., and Sun, S. S. (1995). The composition of the Earth. Chem. Geol.
composition of the Earth’s core. Geophys. Res. Lett. 28 (1), 81–84. doi:10.1029/
120 (3-4), 223–253. doi:10.1016/0009-2541(94)00140-4
2000gl012114
Mel’nik, Y. P. (1972). Thermodynamic parameters of compressed gases and
Li, J., Bergin, E. A., Blake, G. A., Ciesla, F. J., and Hirschmann, M. M. (2021). Earth’s
metamorphic reaction involving water and carbon dioxide. Geochim. Inter. 9, 419–425.
carbon deficit caused by early loss through irreversible sublimation. Sci. Adv. 7 (14),
eabd3632. doi:10.1126/sciadv.abd3632 Mikhail, S., and Sverjensky, D. A. (2014). Nitrogen speciation in upper mantle fluids
and the origin of Earth’s nitrogen-rich atmosphere. Nat. Geosci. 7 (11), 816–819. doi:10.
Li, Y., Dasgupta, R., and Tsuno, K. (2015). The effects of sulfur, silicon, water, and
1038/ngeo2271
oxygen fugacity on carbon solubility and partitioning in Fe-rich alloy and silicate melt
systems at 3 GPa and 1600 °C: implications for core–mantle differentiation and Misener, W., and Schlichting, H. E. (2021). To cool is to keep: residual H/He
degassing of magma oceans and reduced planetary mantles. Earth Planet. Sci. Lett. atmospheres of super-Earths and sub-Neptunes. Mon. Notices R. Astronomical Soc.
415, 54–66. doi:10.1016/j.epsl.2015.01.017 503 (4), 5658–5674. doi:10.1093/mnras/stab895
Li, Y., Dasgupta, R., Tsuno, K., Monteleone, B., and Shimizu, N. (2016b). Carbon and Moore, K., and Cowan, N. B. (2020). Keeping M-Earths habitable in the face of
sulfur budget of the silicate Earth explained by accretion of differentiated planetary atmospheric loss by sequestering water in the mantle. Mon. Notices R. Astronomical Soc.
embryos. Nat. Geosci. 9 (10), 781–785. doi:10.1038/ngeo2801 496 (3), 3786–3795. doi:10.1093/mnras/staa1796
Li, Y., Vočadlo, L., Sun, T., and Brodholt, J. P. (2020). The Earth’s core as a reservoir of Moran, S. E., Stevenson, K. B., Sing, D. K., MacDonald, R. J., Kirk, J., Lustig-Yaeger, J.,
water. Nat. Geosci. 13 (6), 453–458. doi:10.1038/s41561-020-0578-1 et al. (2023). High tide or riptide on the cosmic shoreline? A water-rich atmosphere or
stellar contamination for the warm super-earth gj 486b from JWST observations.
Li, Y., Wiedenbeck, M., Shcheka, S., and Keppler, H. (2013). Nitrogen solubility in
Astrophysical J. Lett. 948 (1), L11. doi:10.3847/2041-8213/accb9c
upper mantle minerals. Earth Planet. Sci. Lett. 377-378, 311–323. doi:10.1016/j.epsl.
2013.07.013 Morard, G., Siebert, J., Andrault, D., Guignot, N., Garbarino, G., Guyot, F., et al.
(2013). The Earth’s core composition from high pressure density measurements of
Li, Y.-F., Marty, B., Shcheka, S., Zimmermann, L., and Keppler, H. (2016a). Nitrogen
liquid iron alloys. Earth Planet. Sci. Lett. 373, 169–178. doi:10.1016/j.epsl.2013.
isotope fractionation during terrestrial core-mantle separation. Geochem. Perspect. Lett.
04.040
2, 138–147. doi:10.7185/geochemlet.1614
Mosenfelder, J. L., Von Der Handt, A., Füri, E., Dalou, C., Hervig, R. L., Rossman, G.
Libourel, G., Marty, B., and Humbert, F. (2003). Nitrogen solubility in basaltic melt.
R., et al. (2019). Nitrogen incorporation in silicates and metals: results from SIMS,
Part I. Effect of oxygen fugacity. Geochimica Cosmochimica Acta 67 (21), 4123–4135.
EPMA, FTIR, and laser-extraction mass spectrometry. Am. Mineralogist J. Earth Planet.
doi:10.1016/s0016-7037(03)00259-x
Mater. 104 (1), 31–46. doi:10.2138/am-2019-6533

Frontiers in Earth Science 21 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Mysen, B. O., and Boettcher, A. L. (1975). Melting of a hydrous mantle: I. Phase Rubie, D. C., Frost, D. J., Mann, U., Asahara, Y., Nimmo, F., Tsuno, K., et al. (2011).
relations of natural peridotite at high pressures and temperatures with controlled Heterogeneous accretion, composition and core–mantle differentiation of the Earth.
activities of water, carbon dioxide, and hydrogen. J. Petrology 16 (1), 520–548. doi:10. Earth Planet. Sci. Lett. 301 (1-2), 31–42. doi:10.1016/j.epsl.2010.11.030
1093/petrology/16.1.520
Rubie, D. C., Laurenz, V., Jacobson, S. A., Morbidelli, A., Palme, H., Vogel, A. K., et al.
Mysen, B. O., and Virgo, D. (1986). Volatiles in silicate melts at high pressure and (2016). Highly siderophile elements were stripped from Earth’s mantle by iron sulfide
temperature: 1. Interaction between OH groups and Si4+, Al3+, Ca2+, Na+ and H+. segregation. Science 353 (6304), 1141–1144. doi:10.1126/science.aaf6919
Chem. Geol. 57 (3-4), 303–331. doi:10.1016/0009-2541(86)90056-2
Rubie, D. C., Melosh, H., Reid, J., Liebske, C., and Righter, K. (2003). Mechanisms of
Mysen, B. O., and Yamashita, S. (2010). Speciation of reduced C–O–H volatiles in metal–silicate equilibration in the terrestrial magma ocean. Earth Planet. Sci. Lett. 205
coexisting fluids and silicate melts determined in-situ to ~1.4GPa and 800°C. (3-4), 239–255. doi:10.1016/s0012-821x(02)01044-0
Geochimica Cosmochimica Acta 74 (15), 4577–4588. doi:10.1016/j.gca.2010.05.004
Rudge, J. F., Kleine, T., and Bourdon, B. (2010). Broad bounds on Earth’s accretion
Namur, O., Charlier, B., Holtz, F., Cartier, C., and McCammon, C. (2016). Sulfur and core formation constrained by geochemical models. Nat. Geosci. 3 (6), 439–443.
solubility in reduced mafic silicate melts: implications for the speciation and distribution doi:10.1038/ngeo872
of sulfur on Mercury. Earth Planet. Sci. Lett. 448, 102–114. doi:10.1016/j.epsl.2016.
Salvador, A., Massol, H., Davaille, A., Marcq, E., Sarda, P., and Chassefière, E. (2017). The
05.024
relative influence of H2O and CO2 on the primitive surface conditions and evolution of rocky
Newman, S., and Lowenstern, J. B. (2002). VolatileCalc: a silicate planets. J. Geophys. Res. Planets 122 (7), 1458–1486. doi:10.1002/2017je005286
melt–H2O–CO2 solution model written in Visual Basic for excel. Comput.
Salvador, A., and Samuel, H. (2023). Convective outgassing efficiency in planetary
Geosciences 28 (5), 597–604. doi:10.1016/s0098-3004(01)00081-4
magma oceans: insights from computational fluid dynamics. Icarus 390, 115265. doi:10.
Nikolaou, A., Katyal, N., Tosi, N., Godolt, M., Grenfell, J. L., and Rauer, H. (2019). 1016/j.icarus.2022.115265
What factors affect the duration and outgassing of the terrestrial magma ocean?
Sasaki, S. (1990). The primary solar-type atmosphere surrounding the accreting
Astrophysical J. 875 (1), 11. doi:10.3847/1538-4357/ab08ed
Earth: H 2 O-induced high surface temperature. Orig. Earth 1990, 195–209.
Nimmo, F., and Schubert, G. (2015). Thermal and compositional evolution of the core.
Schlichting, H. E., and Young, E. D. (2022). Chemical equilibrium between cores,
Elsevier Amsterdam: Core Dynamics, Treatise on Geophysics, 217–241.
mantles, and atmospheres of super-earths and sub-neptunes and implications for their
Nowak, M., and Behrens, H. (1995). The speciation of water in haplogranitic glasses compositions, interiors, and evolution. Planet. Sci. J. 3 (5), 127. doi:10.3847/psj/ac68e6
and melts determined by in situ near-infrared spectroscopy. Geochimica Cosmochimica
Schönbächler, M., Carlson, R. W., Horan, M. F., Mock, T. D., and Hauri, E. H. (2010).
Acta 59 (16), 3445–3450. doi:10.1016/0016-7037(95)00237-t
Heterogeneous accretion and the moderately volatile element budget of earth. Science
O’Brien, D. P., Walsh, K. J., Morbidelli, A., Raymond, S. N., and Mandell, A. M. 328 (5980), 884–887. doi:10.1126/science.1186239
(2014). Water delivery and giant impacts in the ‘Grand Tack’scenario. Icarus 239,
Sharp, Z. D. (2017). Nebular ingassing as a source of volatiles to the Terrestrial
74–84. doi:10.1016/j.icarus.2014.05.009
planets. Chem. Geol. 448, 137–150. doi:10.1016/j.chemgeo.2016.11.018
Ohtani, E., Yurimoto, H., and Seto, S. (1997). Element partitioning between metallic liquid,
Shi, L., Lu, W., Kagoshima, T., Sano, Y., Gao, Z., Du, Z., et al. (2022). Nitrogen isotope
silicate liquid, and lower-mantle minerals: implications for core formation of the Earth. Phys.
evidence for Earth’s heterogeneous accretion of volatiles. Nat. Commun. 13 (1), 4769.
Earth Planet. interiors 100 (1-4), 97–114. doi:10.1016/s0031-9201(96)03234-7
doi:10.1038/s41467-022-32516-5
Okuchi, T. (1997). Hydrogen partitioning into molten iron at high pressure:
Siebert, J., Badro, J., Antonangeli, D., and Ryerson, F. J. (2012). Metal–silicate
implications for Earth’s core. Science 278 (5344), 1781–1784. doi:10.1126/science.
partitioning of Ni and Co in a deep magma ocean. Earth Planet. Sci. Lett. 321-322,
278.5344.1781
189–197. doi:10.1016/j.epsl.2012.01.013
O’Neill, H. S. C., and Mavrogenes, J. A. (2002). The sulfide capacity and the sulfur
Solomatov, V. (2007). “Magma Oceans and primordial mantle differentiation,” in
content at sulfide saturation of silicate melts at 1400degreesC and 1 bar. J. Petrology 43
Treatise on geophysics: evolution of the Earth. Editors G. Schubert and D. Stevenson
(6), 1049–1087. doi:10.1093/petrology/43.6.1049
(United States: Elsevier Ltd).
Otegi, J. F., Bouchy, F., and Helled, R., 2020, Revisited mass-radius relations for
Solomatova, N. V., and Caracas, R. (2019). Pressure-induced coordination
exoplanets below 120 M⊕. Astronomy Astrophysics 634:A43, doi:10.1051/0004-6361/
changes in a pyrolitic silicate melt from ab initio molecular dynamics
201936482
simulations. J. Geophys. Res. Solid Earth 124 (11), 11232–11250. doi:10.1029/
Pahlevan, K., Schaefer, L., Elkins-Tanton, L. T., Desch, S. J., and Buseck, P. R. (2022). 2019jb018238
A primordial atmospheric origin of hydrospheric deuterium enrichment on Mars. Earth
Sossi, P. A., Tollan, P. M., Badro, J., and Bower, D. J. (2023). Solubility of water in
Planet. Sci. Lett. 595, 117772. doi:10.1016/j.epsl.2022.117772
peridotite liquids and the prevalence of steam atmospheres on rocky planets. Earth
Papale, P., Moretti, R., and Barbato, D. (2006). The compositional dependence of the Planet. Sci. Lett. 601, 117894. doi:10.1016/j.epsl.2022.117894
saturation surface of H2O+ CO2 fluids in silicate melts. Chem. Geol. 229 (1-3), 78–95.
Speelmanns, I. M., Schmidt, M. W., and Liebske, C. (2018). Nitrogen solubility
doi:10.1016/j.chemgeo.2006.01.013
in core materials. Geophys. Res. Lett. 45 (15), 7434–7443. doi:10.1029/
Peslier, A. H. (2010). A review of water contents of nominally anhydrous natural 2018gl079130
minerals in the mantles of Earth, Mars and the Moon. J. Volcanol. Geotherm. Res. 197
Stanley, B. D., Hirschmann, M. M., and Withers, A. C. (2014). Solubility of COH
(1-4), 239–258. doi:10.1016/j.jvolgeores.2009.10.006
volatiles in graphite-saturated martian basalts. Geochimica Cosmochimica Acta 129,
Peterson, L. D., Newcombe, M. E., Alexander, C. M. O., Wang, J., Sarafian, A. R., 54–76. doi:10.1016/j.gca.2013.12.013
Bischoff, A., et al. (2023). The H2O content of the ureilite parent body. Geochimica
Steenstra, E. S., Berndt, J., Klemme, S., Rohrbach, A., Bullock, E., and van Westrenen,
Cosmochimica Acta 340, 141–157. doi:10.1016/j.gca.2022.10.036
W. (2020). An experimental assessment of the potential of sulfide saturation of the
Piette, A. A. A., Gao, P., Brugman, K., and Shahar, A. (2023). Rocky planet or water source regions of eucrites and angrites: implications for asteroidal models of core
world? Observability of low-density lava world atmospheres. arXiv preprint arXiv: formation, late accretion and volatile element depletions. Geochimica Cosmochimica
2306.10100. Available at: https://doi.org/10.48550/arXiv.2306.10100. Acta 269, 39–62. doi:10.1016/j.gca.2019.10.006
Righter, K., and Drake, M. J. (1996). Core formation in earth’s moon, Mars, and vesta. Steenstra, E. S., Knibbe, J., and van Westrenen, W. (2016). Constraints on core
Icarus 124 (2), 513–529. doi:10.1006/icar.1996.0227 formation in Vesta from metal–silicate partitioning of siderophile elements. Geochimica
Cosmochimica Acta 177, 48–61. doi:10.1016/j.gca.2016.01.002
Righter, K., and Drake, M. J. (1997). A magma ocean on Vesta: core formation and
petrogenesis of eucrites and diogenites. Meteorit. Planet. Sci. 32 (6), 929–944. doi:10. Steenstra, E. S., and van Westrenen, W. (2020). Geochemical constraints on core-
1111/j.1945-5100.1997.tb01582.x mantle differentiation in Mercury and the aubrite parent body. Icarus 340, 113621.
doi:10.1016/j.icarus.2020.113621
Ringwood, A. E. (1977). Composition of the core and implications for origin of the
Earth. Geochem. J. 11 (3), 111–135. doi:10.2343/geochemj.11.111 Stevenson, D. J. (1981). Models of the Earth’s core. Science 214 (4521), 611–619.
doi:10.1126/science.214.4521.611
Rogers, J. G., and Owen, J. E. (2021). Unveiling the planet population at birth. Mon.
Notices R. Astronomical Soc. 503 (1), 1526–1542. doi:10.1093/mnras/stab529 Stevenson, D. J. (1988). Fluid dynamics of core formation. Top. Conf. Orig. Earth
681, 87.
Rose-Weston, L., Brenan, J. M., Fei, Y., Secco, R. A., and Frost, D. J. (2009). Effect of
pressure, temperature, and oxygen fugacity on the metal-silicate partitioning of Te, Se, Stolper, E. (1982). The speciation of water in silicate melts. Geochimica Cosmochimica
and S: implications for earth differentiation. Geochimica Cosmochimica Acta 73 (15), Acta 46 (12), 2609–2620. doi:10.1016/0016-7037(82)90381-7
4598–4615. doi:10.1016/j.gca.2009.04.028
Suer, T.-A., Siebert, J., Remusat, L., Day, J. M. D., Borensztajn, S., Doisneau, B., et al.
Roskosz, M., Bouhifd, M., Jephcoat, A., Marty, B., and Mysen, B. (2013). Nitrogen (2021). Reconciling metal–silicate partitioning and late accretion in the Earth. Nat.
solubility in molten metal and silicate at high pressure and temperature. Geochimica Commun. 12 (1), 2913–3010. doi:10.1038/s41467-021-23137-5
Cosmochimica Acta 121, 15–28. doi:10.1016/j.gca.2013.07.007
Suer, T.-A., Siebert, J., Remusat, L., Menguy, N., and Fiquet, G. (2017). A sulfur-poor
Rubie, D. C., Nimmo, F., and Melosh, H. J. (2007). Treatise on geophysics. Formation terrestrial core inferred from metal–silicate partitioning experiments. Earth Planet. Sci.
of Earth’s core 9, 51–90. Lett. 469, 84–97. doi:10.1016/j.epsl.2017.04.016

Frontiers in Earth Science 22 frontiersin.org


Suer et al. 10.3389/feart.2023.1159412

Tagawa, S., Sakamoto, N., Hirose, K., Yokoo, S., Hernlund, J., Ohishi, Y., et al. (2021). Wood, B. J. (2008). Accretion and core formation: constraints from metal–silicate
Experimental evidence for hydrogen incorporation into Earth’s core. Nat. Commun. 12 partitioning. Philosophical Trans. R. Soc. A Math. Phys. Eng. Sci. 366 (1883), 4339–4355.
(1), 2588. doi:10.1038/s41467-021-22035-0 doi:10.1098/rsta.2008.0115
Tian, F., Toon, O. B., Pavlov, A. A., and De Sterck, H. (2005). A hydrogen-rich Wood, B. J., Walter, M. J., and Wade, J. (2006). Accretion of the Earth and segregation
early Earth atmosphere. Science 308 (5724), 1014–1017. doi:10.1126/science. of its core. Nature 441 (7095), 825–833. doi:10.1038/nature04763
1106983
Wordsworth, R., and Kreidberg, L. (2022). Atmospheres of rocky exoplanets. Annu.
Tonks, W. B., and Melosh, H. J. (1993). Magma ocean formation due to giant impacts. Rev. Astronomy Astrophysics 60, 159–201. doi:10.1146/annurev-astro-052920-125632
J. Geophys. Res. Planets 98 (E3), 5319–5333. doi:10.1029/92je02726
Wordsworth, R. D. (2016). Atmospheric nitrogen evolution on earth and Venus.
Tsuno, K., Grewal, D. S., and Dasgupta, R. (2018). Core-mantle fractionation of Earth Planet. Sci. Lett. 447, 103–111. doi:10.1016/j.epsl.2016.04.002
carbon in Earth and Mars: the effects of sulfur. Geochimica Cosmochimica Acta 238,
Wu, J., Desch, S. J., Schaefer, L., Elkins-Tanton, L. T., Pahlevan, K., and Buseck, P. R.
477–495. doi:10.1016/j.gca.2018.07.010
(2018). Origin of Earth’s water: chondritic inheritance plus nebular ingassing and storage of
Vander Kaaden, K. E., and McCubbin, F. M. (2015). Exotic crust formation on hydrogen in the core. J. Geophys. Res. Planets 123 (10), 2691–2712. doi:10.1029/2018je005698
Mercury: consequences of a shallow, FeO-poor mantle. J. Geophys. Res. Planets 120 (2),
Xue, X., and Kanzaki, M. (2004). Dissolution mechanisms of water in depolymerized
195–209. doi:10.1002/2014je004733
silicate melts: constraints from 1H and 29Si NMR spectroscopy and ab initio calculations.
Vazan, A., Sari, R., and Kessel, R. (2022). A new perspective on the interiors of ice-rich Geochimica Cosmochimica Acta 68 (24), 5027–5057. doi:10.1016/j.gca.2004.08.016
planets: ice–rock mixture instead of ice on top of rock. Astrophysical J. 926 (2), 150.
Yoshioka, T., Nakashima, D., Nakamura, T., Shcheka, S., and Keppler, H. (2019).
doi:10.3847/1538-4357/ac458c
Carbon solubility in silicate melts in equilibrium with a CO-CO2 gas phase and
Wade, J., and Wood, B. J. (2001). The Earth’s ‘missing’niobium may be in the core. graphite. Geochimica cosmochimica acta 259, 129–143. doi:10.1016/j.gca.2019.06.007
Nature 409 (6816), 75–78. doi:10.1038/35051064
Young, E. D., Shahar, A., and Schlichting, H. E. (2023). Earth shaped by primordial
Wade, J., and Wood, B. J. (2005). Core formation and the oxidation state of the Earth. H2 atmospheres. Nature 616 (7956), 306–311. doi:10.1038/s41586-023-05823-0
Earth Planet. Sci. Lett. 236 (1-2), 78–95. doi:10.1016/j.epsl.2005.05.017
Yuan, L., and Steinle-Neumann, G. (2020). Strong sequestration of hydrogen into the
Wade, J., Wood, B. J., and Tuff, J. (2012). Metal–silicate partitioning of Mo and W at Earth’s core during planetary differentiation. Geophys. Res. Lett. 47 (15),
high pressures and temperatures: evidence for late accretion of sulphur to the Earth. e2020GL088303. doi:10.1029/2020gl088303
Geochimica Cosmochimica Acta 85, 58–74. doi:10.1016/j.gca.2012.01.010
Zahnle, K., Schaefer, L., and Fegley, B. (2010). Earth’s earliest atmospheres. Cold
Wahl, S. M., and Militzer, B. (2015). High-temperature miscibility of iron and rock Spring Harb. Perspect. Biol. 2 (10), a004895. doi:10.1101/cshperspect.a004895
during terrestrial planet formation. Earth Planet. Sci. Lett. 410, 25–33. doi:10.1016/j.
Zahnle, K. J., Kasting, J. F., and Pollack, J. B. (1988). Evolution of a steam atmosphere
epsl.2014.11.014
during earth’s accretion. Icarus 74 (1), 62–97. doi:10.1016/0019-1035(88)90031-0
Walter, M. J., and Cottrell, E. (2013). Assessing uncertainty in geochemical models for
Zahnle, K. J., Lupu, R., Catling, D. C., and Wogan, N. (2020). Creation and evolution
core formation in Earth. Earth Planet. Sci. Lett. 365, 165–176. doi:10.1016/j.epsl.2013.
of impact-generated reduced atmospheres of early Earth. Planet. Sci. J. 1 (1), 11. doi:10.
01.014
3847/psj/ab7e2c
Wang, Z., and Becker, H. (2013). Ratios of S, Se and Te in the silicate Earth require
Zhang, H. L., Cottrell, E., Solheid, P. A., Kelley, K. A., and Hirschmann, M. M. (2018).
a volatile-rich late veneer. Nature 499 (7458), 328–331. doi:10.1038/nature12285
Determination of Fe3+/ΣFe of XANES basaltic glass standards by Mössbauer
Wetzel, D. T., Rutherford, M. J., Jacobsen, S. D., Hauri, E. H., and Saal, A. E. (2013). spectroscopy and its application to the oxidation state of iron in MORB. Chem.
Degassing of reduced carbon from planetary basalts. Proc. Natl. Acad. Sci. 110 (20), Geol. 479, 166–175. doi:10.1016/j.chemgeo.2018.01.006
8010–8013. doi:10.1073/pnas.1219266110
Zhang, M., and Li, Y. (2021). Breaking of Henry’s law for sulfide liquid–basaltic melt
Wolf, A. S., Jäggi, N., Sossi, P. A., and Bower, D. J. (2022). VapoRock: partitioning of Pt and Pd. Nat. Commun. 12 (1), 5994. doi:10.1038/s41467-021-26311-x
thermodynamics of vaporized silicate melts for modeling volcanic outgassing and
Zieba, S., Kreidberg, L., Ducrot, E., Gillon, M., Morley, C., Schaefer, L., et al. (2023).
magma ocean atmospheres. arXiv preprint arXiv:2208.09582. Available at: https://doi.
No thick carbon dioxide atmosphere on the rocky exoplanet TRAPPIST-1 c. Nature
org/10.48550/arXiv.2208.09582.
620, 746–749. doi:10.1038/s41586-023-06232-z

Frontiers in Earth Science 23 frontiersin.org

You might also like